Statistical Aspects of Clinical Trials PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 352

STATISTI(AL ASPKTS

OF THE DlSlGN
AND ANAlYSlS Of
(lINI(A1 TRIALS
STATISTICAL ASPECTS
OF THE DESIGN
AND ANAlYSlS OF
(llNKA1 TRlAlS

I
Brian 5. Everitt
lnrtitute of Piychiatry, London

Andrew Picklei
Univetrity of Monchertet

Imperial College Press


Published by
Imperial College Press
57 Shelton Street
Covent Garden
London WC2H 9HE

Distributed by
World Scientific Publishing Co. Re. Ltd.
P 0 Box 128, Farrer Road, Singapore 912805
USA ofice: Suite lB, 1060 Main Street, River Edge, NJ 07661
UKofJice: 57 Shelton Street, Covent Garden, London WC2H 9HE

British Library Cataloguing-in-PublicationData


A catalogue record for this book is available from the British Library.

STATISTICAL ASPECTS OF THE DESIGN AND ANALYSIS OF CLINICAL TRIALS


Copyright 0 1999 by Imperial College Press
All rights reserved. This book, or parts thereof; may not be reproduced in any form or by any means,
electronic or mechanical, including photocopying, recording or any information storage and retrieval
system now known or to be invented, without written permission from the Publisher.

For photocopying of material in this volume, please pay a copying fee through the Copyright
Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to
photocopy is not required from the publisher.

ISBN 1-86094-153-2

This book is printed on acid-free paper.

Printed in Singapore by Regal Press (S) Re. Ltd.


PREFACE
According to Sir David Cox, the randomised controlled clinical trial
is perhaps the outstanding contribution of statistics to 20th century
medical research. Nowadays about 8000 such trials are undertaken
annually in all areas of medicine from the treatment of acne to the
prevention of cancer. Although the vast majority of these trials take
place away from the glare of public interest, some deal with issues
that are controversial enough to make even the popular press; an
obvious example is the use of AZT for the treatment of AIDS.
There are many excellent books available which give comprehen-
sive accounts of how clinical trials should be carried out and organ-
ised. Our aim is somewhat different; we attempt to give relatively
concise descriptions of the more statistical aspects of the design and
analysis of clinical trials, particularly those methods developed over
the last decade or so. Topics discussed in this text include randomi-
sation, interim analyses, sample size determination, the analysis of
longitudinal data, Bayesian methods, survival analysis and meta-
analysis. Many examples are included alongside some of the neces-
sary technical material, the more difficult parts of which are confined
to tables. An Appendix gives details of relevant software. We hope
that our book will be useful to medical statisticians and others faced
with the often difficult problems of designing and analysing clinical
trials.
Our thanks are due to Dr Sophia Rabe-Hesketh and Dr Sabine
Landau for reading the text and making many helpful suggestions,
to Professor Elizabeth Kuipers for allowing us to use the economic
data from her CBT trial in Chapter 4 and to Mrs Harriet Meteyard
for help in compiling the references.

Brian S. Everitt and Andrew Pickles


London, 1999

V
CONTENTS

PREFACE V

CHAPTER 1. An Introduction to Clinical Trials 1


1.1 INTRODUCTION 1
1.2 A BRIEF HISTORY OF
CLINICAL TRIALS 2
1.3 TYPES O F CLINICAL TRIAL 5
1.4 ETHICS OF CLINICAL TRIALS 9
1.5 CLINICAL TRIAL PROTOCOLS 16
1.6 SUMMARY 16
CHAPTER 2. Treatment Allocation, the Size of
Trials and Reporting Results 18
2.1 INTRODUCTION 18
2.2 TREATMENT ASSIGNMENT
METHODS 18
2.2.1 Simple Randomisation 21
2.2.2 Blocked Randomisation 22
2.2.3 Stratified Randomisation 24
2.2.4 Minimisation Method 26
2.2.5 Unequal Randomisation 27
2.3 THE SIZE OF A CLINICAL TRIAL 29
2.4 REPORTING RESULTS 33
2.5 SUMMARY 35

vii
viii Design and Analysis of Clinical Trials

CHAPTER 3. Monitoring Trial Progress: Outcome


Measures, Compliance, Dropouts
and Interim Analyses 37
3.1 INTRODUCTION 37
3.2 OUTCOME MEASURES 39
3.3 COMPLIANCE, DROPOUTS AND
INTENTION-TO-TREAT 41
3.3.1 Int,ention-tc-Treat 45
3.3.2 A Taxonomy of Dropouts 50
3.4 INTERIM ANALYSES IN
CLINICAL TRIALS 54
3.4.1 Group Sequential Procedures
- A Bayesian Perspective 65
3.5 SUMMARY 67
CHAPTER 4. Basic Analyses of Clinical Trials,
the Generalized Linear Model and
the Economic Evaluation of Trials 69
4.1 INTRODUCTION 69
4.2 A BRIEF REVIEW OF EL~SIC
STATISTICS 70
4.3 GENERALIZED LINEAR
MODELS 77
4.3.1 Models for Counts: the
Treatment of Familial
Adenomatous Poiyposis (FAP)
wit.h a Non-Steroidal
Anti-Inflammatory Drug 83
4.3.2 Ordinal Response Models:
A Clinical Trial in
Lymphoma 89
4.3.2.1 Models with parallel
linear predictors 89
4.3.2.2 Non-parallel models 90
4.3.2.3 Model estimation 91
Contents ix

4.4 MULTIPLE EIYDPOINTS 93


4.5 ECONOMIC EVALUATION
O F TRTALS 97
4.5.1 Measuring Costs 97
4.5.2 Analysis of Cost Data 99
4.5.3 Cost-Benefit Analysis 101
4.5.4 An Example: Cognitive
Behavioral Therapy 101
4.6 SUMMARY 106

CHAPTER 5 . Simple Approaches to the Analysis of


Longitudinal Data from Clinical Trials 108
5.1 INTRODUCTION 108
5.2 GRAPHICAL METHODS FOR
DISPLAYING LONGITUDINAL
DATA 109
5.3 TIME-BY-TIME ilNALYSIS
O F LONGITUDINAL DATA 118
5.4 RESPONSE FEATURE ANALYSIS
O F LONGITUDINAL DATA 120
5.4.1 Choosing Summary Measures 121
5.4.2 Incorporating Pre-Randomisation,
Baseline Measures int,o the
Response Approach 129
5.4.3 Response Feature Analysis
when the Response Variable
is Binary 134
5.5 SUMMARY 142

CHAPTER 6 . Multivariate Normal Regression Models for


Longitudinal Data from Clinical Trials 143
6.1 INTRODUCTION 143
6.2 SOME GENERAL COMMENTS
ABOUT MODELS FOR
LONGITUDINAL DATA 144
X Design and Analysis of Clinical Trials

6.3 MISSING VALUES AND


LIKELIHO OD-B ASED INFERENCE
FOR LONGITUDINAL DATA 147
6.4 MAXIMUM LIKELIHOOD
ESTIMATION ASSUMING
NORMALITY 149
6.5 NUMERICAL EXAMPLES 151
6.5.1 Oestrogen Patches in the
Treatment of Post*-N atal
Depression 151
6.5.2 Post-Surgical Recovery in
Young Children 161
6.6 ANOVA AND MANOVA
APPROACHES TO THE ANALYSIS
O F LONGITUDINAL DATA4 167
6.7 INFORMATIVE DROP OUTS 168
6.8 OTHER METHODS FOR THE
ANALYSIS OF LONGITUDINAL
DATA 173
6.9 SUMMA4RY 175
CHAPTER 7. Models for Non-Normal Longitudinal
Data from Clinical Trials 178
7.1 INTRODUCTION 178
7.2 MARGINAL MODELS 179
7.2.1 Marginal Modelling using
Generalised Estimating
Equations 183
7.3 RANDOM EFFECTS MODELS 193
7.4 TRANSITION MODELS 197
7.5 A COMPARISON O F METHODS 201
7.6 MISSING DATA AND WEIGHTED
ESTIMATING EQUATIONS 202
Contents xi

7.6.1
Missing Data in a Longitudinal
Clinical Trial with an
Ordinal Response 203
7.7 SUMMARY 206
CHAPTER 8 . Survival Analysis 208
8.1 INTRODUCTIOK 208
8.2 THE SURVIVOR FUNCTION AND
THE HAZARD FUNCTION 209
8.2.1 Survivor Function 2 09
8.2.2 The Hazard Function 219
8.3 REGRESSION MODELS FOR
SURVIVAL DATA 220
8.4 ACCELERATED FAILURE TIME
MODELS 222
8.5 PROPORTIONAL HAZARDS MODEL 225
8.5.1 Proportional Hazards 225
8.5.2 The Semi-parametric
Proportional Hazard
or Cox Model 226
5.5.3 Cox Model Example 229
8.5.4 Checking the Specification
of a Cox Model 232
8.5.5 Goodness-of-fit Tests 240
8.5.6 Influence 241
8.6 TI ME-17ARYlh’GCOVARIATE S 242
8.6.1 Modelling with Time-Varying
Covariates 242
8.6.2 The Probiem of Int,ernal or
Endogeneous Covariates 243
8.6.3 Treatment Waiting Times 244
8.7 STRATIFICATION, MATCHING
AND CLUSTER SAMPLING 245
xii Design and Analysis of Clinical Trials

8.7.1 Stratification 245


8.7.2 Matching 246
8.7.3 Cluster Sampling 247
8.8 CENSORING AND COMPETING
RISKS 248
8.9 AUXILIARY VARIABLES 249
8.10 FRAILTY MODELS 250
8.11 INTERIM ANALYSIS FOR
SURVIVAL TIME DATA 251
8.11.1 An Example of the Application
of an Interim Analysis
Procedure for Survival Data 251
8.12 DESCRIBING THE RESULTS O F
A SURVIVAL ANALYSIS OF A
CLINICAL TRIAL 252
8.13 SUMMARY 253
CHAPTER 9. Bayesian Methods 255
9.1 INTRODUCTION 255
9.2 BAYESIAN ESTIMATION 257
9.3 MARKOV CHAIN MONTE
CARL0 (MCMC) 259
9.4 ADVANTAGES OF THE
MCMC APPROACH 260
9.4.1 Model Parametrisation and
Scientific Inference 261
9.4.2 Missing Data 262
9.4.3 Prior Distributions as
a Numerical Device 262
9.4.4 Alternative Models and
Model Selection 263
9.5 A NUMERICAL EXAMPLE OF
USING BAYESIAN ESTIMATION 264
9.6 INFORMATIVE PRIORS 268
Contents
...
Xlll

9.7 PRIORS AND STOPPING RULES


FOR TRIAL MONITORING 268
9.7.1 Clinical Demands 269
9.7.2 Clinical Beliefs 271
9.8 MONITORING 272
9.8.1 The Prior as a Handicap 272
9.8.2 Recognising Heterogeneity in
Priors: Enthusiasts and Sceptics 273
9.9 SAMPLE SIZE ESTIMATION FOR
EARLY PHASE TRIALS 275
9.10 REPORTING O F RESULTS 278
9.11 SUMMARY 279
CHAPTER 10. Meta-Analysis 280
10.1 INTRODUCTION 280
10.2 SELECTION O F STUDIES
- WHAT T O INCLUDE 282
10.3 THE STATISTICS OF
META-ANALYSIS 290
10.4 TWO EXAMPLES O F THE
APPLICATION OF
META-ANALYSIS 297
10.4.1 The Comparison of Sodium
Monofluorophosphate (SMFP)
and Sodium Fluoride (NaF)
Toothpastes in the Prevention
of Caries Development 297
10.4.2 The Effectiveness of Aspirin
in Preventing Death after a
Myocardial Infarction 297
10.5 BAYESIAN META-ANALYSIS 300
xiv Design and Analysis of Clinical R i a l s

10.6 META-ANALYSIS IN MORE


COMPLEX SETTINGS 302
10.7 SUMMARY 303
APPENDIX A. Software 305
REFERENCES 311
INDEX 333
CHAPTER
1

An Introduction t o
Clinical Trials

1.1. INTRODUCTION

Avicenna, an Arabian physician and philospher (98&1037), in his


encyclopedic Canon of Medicine, set down seven rules t o evaluate
the effect of drugs on diseases. He suggested that a remedy should
be used in its natural state, with uncomplicated disease, and should
be observed in two ‘contrary types of disease.’ His Canon also sug-
gested that the time of action and reproducibility of the treatment
effect should be studied (Crombie, 1952; Meinert, 1986).
But for several centuries Avicenna’s advice appears to have been
largely ignored, with most ideas affecting choice of treatment depend-
ing largely on serendipity rather then planned experiments. Only in
recent years (although see next section) has it become widely recog-
nised that properly conducted clinical trials, which follow the prin-
ciple of scientific experimentation provide the only reliable basis for
evaluating the efficacy and safety of new treatments.
And just what constitutes a clinical trial? There are several pos-
sible definitions, but for our purposes the term will be used for any
form of planned experiment designed to assess the most appropri-
ate treatment of future patients with a particular medical condition,
where the outcome in a group of patients treated with the test treat-
ment are compared with those observed in a similar group of pa-
tients receiving a control treatment, and patients in both groups are

1
2 Design and Analysis of Clinical Trials

enrolled, treated and followed over the same time period. The groups
may be established through randomisation or some other method of
assignment. The outcome measure may be the result of a laboratory
test, a quality of life assessment, a rating of some characteristic or,
in some cases, the death of a patient.
As a consequence of this somewhat restricted definition, compar-
ative studies involving animals, or studies that are carried out in vitro
using biological substances from man do not qualify as clinical trials.
The definition also rules out detailed consideration of investigations
involving historical controls.

1.2. A BRIEF HISTORY OF CLINICAL


TRIALS

It is almost de rigeur in books on clinical trials to include a section


tracing their history. Our book is no exception! Table 1.1 (taken
from Meinert, 1986) lists some important dates in the development
of such trials, the first of which relates to the often described exper-
iment of James Lind carried out in 1747 while at sea on board the
Salisbury. Bradford Hill (1962) gives the following quotation from
Lind’s account.

On the 20th May 1747, I took twelve patients in the scurvy, on board
the Salisbury at sea. Their cases were as similar as I could have them. They
all in general had putrid gums, the spots and lassitude, with weakness of
their knees. They lay together in one place, being a proper apartment for
the sick in the fore-hold; and had one diet in common to all, viz. water-
gruel sweetened with sugar in the morning; fresh mutton broth often times
for dinner; at other times puddings, boiled biscuit with sugar etc. And for
supper, barley and raisins, rice and currants, sago and wine, or the like.
Two of these were ordered each a quart of cider a day. Two others took
twenty-five gutts of elixir vitriol three times a day, upon an empty stomach;
using a gargle strongly acidulated with it for their mouths. Two others took
two spoonfuls of vinegar three times a day, upon an empty stomach: having
their gruels and their other food well acidulated with it, as also the gargle
An Introduction to Clinical Trials 3

Table 1.1. Historical Events in the Development of Clinical Trials.

Date Author Event

1747 Lind Experiment with untreated control group (Lind, 1753)


1799 Haygarth Use of sham procedure (Haygarth, 1800)
1800 Waterhouse U.S.-based smallpox trial (Waterhouse, 1800, 1802)
1863 Gull Use of placebo treatment (Sutton, 1865)
1923 Fisher Application of randomisation to experimentation
(Fisher and MacKenzie. 1923)
1931 - Special committee on clinical trial created by the
Medical Research Council of Great Britain
(Medical Research Council, 1931)
1931 Amberson Random allocation of treatment to groups of
patients (Amberson et al., 1931)
1937 - Start of NIH grant support with creation of the
National Cancer Institute (National Institutes of
Health. 1981b)
1944 ~

Publication of multicenter trial on treatment for


common cold (Patulin Clinical Trials Committee, 1944)
1946 - Promulgation of Nurernberg Code for Human
Experimentation (Curran and Shapiro, 1970)
1962 Hill Publication of book on clinical trials (Hill, 1962)
1962 Kefauver Amendments to the Food, Drug and Cosmetic Act
and Harris of 1938 (United States Congress, 1962)
1966 - Publication of U.S. Public Health Service
regulations leading to creation of Institutional
Review Boards for research involving humans
(Levine, 1981)
1967 Chalmers Structure for separating the treatment monitoring
and treatment administration process
(Coronary Drug Project Research Group, 1973a)
1979 - Establishment of Society for Clinical Trials
(Society for Clinical Trials, Inc., 1980)
1980 - First issue of Controlled Clinical Trials

(Taken with permission from Meinert, 1986.)


4 Design and Analysis of Clinical Trials

for their mouths. Two of the worst patients, with the tendons in the ham
rigid (a symptom none of the rest had) were put under a course of sea-water.
Of this they drank half a pint every day, and sometimes more or less as
it operated, by way of a gentle physic. Two others had each two oranges
and one lemon given them every day. These they eat with greediness,
at different times, upon an empty stomach. They continued but six days
under this course, having consumed the quantity that could be spared. The
two remaining patients, took the bigness of a nutmeg three times a day of
an electuary recommended by a hospital-surgeon, made of garlic, mustard-
feed, rad. raphan. balsam of Peru, and gum myrr; using for common drink
barley water well acidulated with tamarinds; by a decoction of which, with
the addition of cremor tartar, they were greatly purged three or four times
during the course. The consequence was, that the most sudden and visible
good effects were perceived from the use of the oranges and lemons; one of
those who had taken them, being at the end of six days fit for duty. The
spots mere not indeed at that time quite off his body, nor his gums sound;
but without any other medicine, than a gargle of elixir vitriol, he became
quite healthy before we came into Plymouth. which was OR the 16th June.
The other was the best recovered of any in his condition; and being now
deemed pretty well, was appointed nurse to the rest of the sick.

In spite of the relative clear-cut nature of his findings, Lind still


advised that the best treatment for scurvy involved placing stricken
patients in ‘pure dry air.’ No doubt the reluctance t o accept oranges
and lemons as treatment for t h e disease had something t o d o with
their expense compared t o the ‘dry air’ treatment. In fact it was a
further 40 years before the British Navy supported lemon juice for
the crews of its ships at sea; once again the question of cost quickly
became a n issue with lemons being substituted by limes, condemming
the British sailor t o be referred to for the next two hundred years as
‘limeys’.
Most of the early experiments involved arbitrary, nonsystematic
schemes for assigning patients to treatments, such as that described
by Lind. T h e concept of randomisation a s a method for treatment
assignment was first introduced by Fisher and the first trial with
A n Introduction to Clinical Trials 5

a properly randomised control group was for streptomycin in the


treatment of pulmonary tuberculosis (see Medical Research Council,
1948, and Armitage, 1983). But not all clinicians were convinced
of the need for such trials - the following is taken from a letter
published in a medical journal of the day, attacking a proposed trial
for the treatment of depression:

There is no psychiatric illness in which bedside knowledge and long


clincal experience pays better dividends; and we are never going to learn
about how to treat depressions properly from double blind sampling in an
MRC statistician’s office.

Since World War 11, the clinical triaI has evolved into a standard
procedure in the evaluation of new drugs. Its features include the use
of a control group of patients that do not receive the experimental
treatment, the random allocation of patients to the experimental or
control group, and the use of blind or masked assessment so that
neither the researchers nor the patients know which patients are in
either group at the time the study is conducted. The clinical trial
nicely illustrates the desire of modern democratic society to justify its
medical choices on the basis of the objectivity inherent in statistical
and quantitative data.

1.3. TYPES OF CLINICAL TRIAL


Clinical trials can take a variety of different forms. All however
are prospective with observations being made over a period of time
after treatment allocation. Perhaps the most common design for
a clinica.1 trial is the fixed sample size parallel groups design with
random allocation of patients to treatment, rather than some larger
randomisation unit such as famiIy, hospital, ward, community, etc.
One problem with such a design occurs when patients vary so much
in their initial disease state and in their response t.0 therapy that
large numbers of patients may be needed to estimate reliably t.he
magnitude of any treatment difference. A more precise treatment
6 Design and Analysis of Clinical Rials

comparison might be achieved by using a crowover design in which


each patient receives more than one treatment. A simple example is
the 2 x 2 cross-over design in which one group of patients receive
two treatments, A and B, in the order AB, another group in the
order BA, with patients being randomly allocated t o the two groups.
Clearly such a design is only suitable for chronic conditions in which
there is the limited objective of studying the patient’s response to
relatively short periods of therapy. The design and analysis of cross-
over trials is more than adequately dealt with in Jones and Kenward
(1989) and Senn (1993), and so will not be considered in any detail in
this text.
The majority of randomised, placebo-controlled clinical trials
have focussed on one drug at a time although this does not match
up with clinical practice where it is rarely sufficient to consider only
a single treatment for a condition. Questions about the effects of
combinations of treatments can never be resolved by the simple par-
allel groups design in which an active treatment is compared with
a placebo; consequently, some investigators have proposed factorial
designs in which several treatments are considered simultaneously.
Lubsen and Pocock (1994), for example, describe a trial in which
patients were simultaneously randomised to each of three active
treatments or their respective controls in a 2 x 2 x 2 factorial arrange-
ment. The claim made for the trial is that it provides three answers
for the price of one (see Collins, 1993). As Lubsen and Pocock point
out, this claim is only justified if it can safely be assumed that there
is no evidence of any interaction between the three treatments. Lack
of interaction implies that the effect of the treatments are additive on
some particular scale expressing the effects of each treatment. Lub-
sen and Pocock are sceptical about whether interactions can often be
dismissed a priori; if they cannot, then factorial designs will require
larger sample sizes to achieve the same power as a parallel groups
design. Their conclusion is that such designs are most appropriate
for assessing therapeutic combinations when possible interactions are
actually of primary interest. Some consideration of such studies is
given in Holtzmann (1987) and Berry (1990).
A n Introduction to Clinical m a 2 s 7

The pharmaceutical industry uses a well-established taxonomy


of clinical trials involving drug therapy, in which the categories can,
according to Pocock (1983), be described as follows:

Phase I Trials: Clinical Pharmacoloy and Toxicity

These first experiments in man are primarily concerned with drug


safety, not efficacy, and hence are usually performed on healthy, hu-
man volunteers, often pharmaceutical company employees. The first
objective is to determine an acceptable single drug dosage (i.e. how
much drug can be given without causing serious side-effects). Such
information is often obtained from dose-escalation experiments,
whereby a volunteer is subjected to increasing doses of the drug ac-
cording to a predetermined schedule. Phase I will also include studies
of drug metabolism and bioavailability a.nd later, studies of multiple
doses will be undertaken to determine appropriate dose schedules for
use in phase 11. After studies in normal volunteers, the initial trials in
patients will also be of phase I type. Typically, phase I studies might
require a total of around 20-80 subjects or patients. The general aim
of such studies is to provide a relatively clear picture of a drug, but
one that will require refinement. during phases I1 and 111.

Phase 11 Trials: Initial Clinical Investigation for Treatment Effect

These are fairly small-scale investigations into the effectiveness


and safety of a drug, and require close monitoring of each patient.
Phase I1 trials can sometimes be set up as a screening process to
select out those relatively few drugs of genuine potential from the
larger number of drugs which are inactive or over-toxic: so that the
chosen drugs may proceed t o phase I11 trials. Seldom will phase I1
go beyond 100-200 patients on a drug. The primary goals of phase I1
trials are:
to identify accurately the patient population that can benefit
from the drug,
8 Design and Analysis of Clinical Trials

0 to verify and estimate the effectiveness of the dosing regimen


determined in phase I.

Phase III Trials: Full-scale Evaluation of Treatment

After a drug is shown t.o be reasonably effective, it is essential


to compare it with the current standard treatment(s) for the same
condition in a large trial involving a substantial number of patients.
To some people t,he term ‘clinical trial’ is synonymous with such a
full-scale phase 111 trial, which is the most rigorous and extensive
type of scientific clinical investigation of a new treatment.

Phase IV Trials: Postmarketing Surveil lance

After the research programme leading to a drug being approved


for marketing, there remain substantial enquiries still to be under-
taken as regards monitoring for adverse effects and additional large-
scale, long-term studies of morbidity and mortality.

This book will be largely concerned with phase I11 trials. In or-
der to accumulate enough patients in a time short enough to make a
trial viable. many such trials will involve recruiting patients at more
than a single centre (for example, a clinic, a hospital, etc.); they
will be multicentre trials. The principal advantage of carrying out
a multicentre trial is that patient accrual is much quicker so that
the trial can be made larger and the planned number of patients can
be achieved more quickly. The end-result should be that a multicen-
tre trial reaches more reliable conclusions at a faster rate, so that
overall progress in the treatment of a given disease is enhanced.
Recommendations over the appropriate number of centres varies;
on the one hand, rate of patient acquisition may be completely inade-
quate when dealing with a smalI number of centres, but with a large
number (20 or more) potential practical problems (see Table 1.2)
may quickly outweigh benefits. There also be other problems involv-
ing the analysas of multi-centre trials. It is likely, for example, that
An Introduction to Clinical Trials 9

Table 1.2. Potential Problems with Multicentre Trials.


The planning and administration of any multicentre trial is consid-
erably more complex than in a single centre,
Multicentre trials are very expensive to run,
Ensuring that all centres follow the study protocol may be difficult,
Consistency of measurements across centres needs very careful
at tent ion,
Motivating all participants in a large multicentre trial may be
difficult,
Lack of clear leadership may lead to a degeneration in the quality
of a multicentre trial.

the true treatment effect will not b e identical at each centre. Conse-
quently there may be some degree of treatment-by-centre interaction
and various methods have been suggested for dealing with this possi-
bility. Details are available in Jones et al. (1998), Gould (1998) and
Senn (1998).

1.4. ETHICS OF CLINICAL TRIALS


Since the time of Hippocrates, Western physicians have taken a n
oath in which they swear t o protect their patients ‘from whatever
is deleterious and mischievous.’ Unfortunately such an oath has not
managed t o stop many damaging therapies being given or t o lessen
the persistence of barbarous practices such as copious blood-letting.
Even the most powerful members of society were vulnerable t o the
ill-informed, if well-intentioned physician, as the following account of
the treatment of the dying Charles I1 demonstrates:

At eight o’clock on Monday morning of February 2, 1685, King Charles


I1 of England was being shaved in his bedroom. With a sudden cry he fell
backward and had a violent convulsion. He became unconscious, rallied
once or twice, and after a few days, died. Doctor Scarburgh, one of the
twelve or fourteen physicians called to treat the stricken king, recorded the
efforts made to cure the patient. As the first step in treatment the king was
bled to the extent of a pint from a vein in his right arm. Next his shoulder
10 Design and Analysis of Clinical Trials

was cut into and the incised area was ‘cupped’ t o suck out an additional
eight ounces of blood. After this, the drugging began. An emetic and
purgative were administered, and soon after a second purgative. This was
followed by an enema containing antimony, sacred bitters, rock salt, mallow
leaves, violets, beetroot, camomile flowers, fennel seed, linseed, cinnamon,
cardamom seed, saphron, cochineal, and aloes. The enema was repeated in
two hours and a purgative given. The king’s head was shaved and a blister
raised on his scalp. A sneezing powder of hellebore root was administered
and also a powder of cowslip flowers ‘to strengthen his brain.’ The cathar-
tics were repeated at frequent intervals and interspersed with a soothing
drink composed of barley water, liquorice, and sweet almond. Likewise
white wine, absinthe, and anise were given, as also were extracts of thistle
leaves, mint, rue, and angelica. For external treatment a plaster of Bur-
gundy pitch and pigeon dung was applied to the king’s feet. The bleeding
and purging continued, and to the medicaments were added melon seeds,
manna, slippery elm, black cherry water, a n extract of flowers of lime, lily
of the valley, peony, lavender, and dissolved pearls. Later came gentian
root, nutmeg, quinine and cloves. The king’s condition did not improve,
indeed it grew worse, and in the emergency forty drops of extract of human
skull were administered to allay convulsions. A rallying dose of Raleigh’s
antidote was forced down the king’s throat; this antidote contained an enor-
mous number of herbs and animal extracts. Finally bezoar stone was given.
“Then”, said Scarburgh, “Alas! after an ill-fated night his serene majesty’s
strength seemed exhausted to such a degree that the whole assembly of
physicians lost all hope and became despondent; still so as not to appear to
fail in doing their duty in any detail, they brought into play the most active
cordial.’’ As a sort of grand summary to this pharmaceutical debauch, a
mixture of Raleigh’s antidote, pearl julep, and ammonia was forced down
the throat of the dying king.

Ethical issues in medicine in general a n d clinical trials in particu-


lar a r e clearly of great importance and present a potential minefield
especially for two statisticians more involved and perhaps more inter-
ested in the pragmatic problems of t h e analysis of the data generated
in such trials. Nonetheless, along with all staff involved in trials, t h e
statistician must share in the general responsibility for the ethical
An Introduction to Clinical Trials 11

conduct of a trial. And there are in addition some areas of trial con-
duct where the statistician needs to take particular responsibility for
ensuring that both the proposed and actual conduct of the trial are
appropriate.
A central ethical issue often identified with clinical trials is t,hat
of randomisation. Randomised controlled trials are now widely used
in medical research. Two recent examples from the many trials un-
dertaken each year include:

0 A multicentre study of a low-protein diet on the progression


of chronic renal failure in children (Wingen et aZ; 1997),
0 A study of immunotherapy for asthma in allergic children (Ad-
kinson Jr. et d.,1997).

Random allocation gives all subjects the same chance of receiving


each possible treatment (although see Chapter 2). Randomisation
serves several purposes; it provides an important method of allocat-
ing patients to treatments free from personal biases and it ensures a
firm basis for the application of significance tests and most of the
rest of the statistical methodology likely t o be used in assessing
the results of the trial. Most importantly, randomisation distributes
the effects of concomitant variables, both measured and unobserved
(and possibly unknown)! in a chance, and therefore, impartial fashion
amongst the groups t o be compared. In this way, random allocation
ensures a lack of bias, making the interpretation of an observed group
difference largely unambiguous - its cause is very likely to be the
different treatments received by the different groups.
Unfortunately, however, the idea that patients should be ran-
domly assigned to treatments is often not appealing to many clini-
cians nor t o many of the individuals who are prospective participants
in a trial. The reasons for their concern are not difficult to identify.
The clinician faced with the responsibility of restoring the patient
to health and suspecting that any new treatment is likely to have
advantages over the old, may be unhappy that many patients will
be receiving, in her view, the less valuable treatment. The patient
12 Design and Analysis of Clinical Trials

being recruited for a trial, having been made aware of the randomi-
sation component, might be troubled by the possibility of receiving
an ‘inferior’ treatment.
Few clinicians would argue against, the need for the voluntary
consent of subjects being asked to take part in a trial, but the amount
of information given in obtaining such consent might be a matter for
less agreement. Most clinicians would accept that the subject must
be allowed to know about the randomisation aspect of the trial, but
how many would want to go as far as Berry (1993) in advising the
subject along the following lines?

I would like you to yart,icipate in a randomised trial. We will in effect


Rip a coin and give you therapy A if the coin comes up heads and therapy
B if it comes up tails. Neither you or I will know what therapy you receive
unless problems develop. [After presenting information about the therapies
and their possible side-effects:] KO one really knows what therapy is better
and that is why we’re conducting this trial. However, we have had some
experience with both therapies, including experience in the current trial.
The available data suggest that you will live an average of five months
longer on A than on B. But there is substantial variability in the data, and
many people who have received B have lived longer than some patients on
A. If I were you I would prefer A. My probability h a t you live longer on
A is 25 per cent.
Your participation in this trial will help us treat other patients with this
disease. so I ask you in their name. But if you choose not to participate,
you will receive whichever therapy you choose, including A or B.

Berry’s suggestion as to how to inform subjects considering taking


part in a clinical trial highlights the main ethical problem in such a
study, namely the possible conflict between trying to ensure that
each individual patient receives the treatment most beneficial for
his/her condition, and evaluating competing therapies as efficiently
as possible so that all future patients might benefit from the superior
treatment. The great dilemma of clinical trials is that if each patient
is treated as well as possible, patients as a whole are not. Lellouch
An Introduction to Clinical Trials 13

and Schwartz (1971) refer to the problem as competition between


individual and collective ethics. Pocock (1983) suggests that each
clinical trial requires a balance between the two. The prime motiva-
tion for conducting a trial involves future patients, but individuals
involved in the trial have to be given as much attention as possible
without the trial’s validity being destroyed. Naturally the clinician’s
responsibility to patients during the course of a trial are clear; if the
patient’s condition deteriorates, the ethical obligation must always
and entirely outweigh any experimental requirements. This obliga-
tion means that whenever a physician thinks that the interest of a
patient are at stake, she must be allowed to treat the patient as she
sees fit. This is an absolutely essential requirement for an ethically
conducted trial, no matter what complications it may introduce into
the final analysis of the resulting data.
Clearly the ethical issues will be of greater concern in trials where
the condition being treated is extremeIy serious, possibly even life
threatening, than when it is more mild. The problems that can
arise in the former situation are well illustrated by the history of
the trials of AZT a s a therapy for AIDS. When such trials were first
announced there was a large, vocal lobby against testing the drug
in a controlled clinical trial where necessarily some patients would
receive an ‘inferior treatment’. Later, however, when the severity of
some side effects was identified and the long term effectiveness of the
drug in doubt, an equally vocal lobby called for AZT treatment to
be abandoned. Expanding networks of ‘support groups’ makes these
problems increasingly likely.
If randomisation is the first priority in an acceptable clinical trial,
blinding comes a close second. The fundamental idea of blinding is
that the trial patients, the people involved with their management
and those collecting clinical data from studies, should not be influ-
enced by knowledge of the assigned treatment. Blinding is needed t o
prevent the possibility of bias arising from the patient, the physician
and in evaluation. There are a number of levels of blinding of which
the two most important are:
14 Design and Analysis of Clinical Dials

0 SzngEe-blind: Usually used for the situation in which the pa-


tient is unaware of which treatment he or she is receiving.
0 Double-blind: Here both the patient and the investigator are
kept blind to the patient’s treatment. For many trials this is
the arrangement of choice.

In drug trials blinding is usually relatively easy t o arrange but


the blinding of physical treatments, for example, surgical procedures,
is often more difficult.
The randomised double-blind controlled trial is the ‘gold-
standard’ against which to judge the quality of clinical trials in gen-
eral. But such trials are still misunderstood by many clinicians and
questions about whether or not they are ethical persist. One of the
problems identified by Bracken (1987), is that doctors are frequently
reluctant to accept their uncertainty about much of what they prac-
tice. Bracken concludes that when doctors are able t o admit t o them-
selves and their patients uncertainty about the best action, then no
conflict exists between the roles of the doctor and the statistician. In
such circumstances it cannot be less ethical to choose a. treatment by
random allocation within a controlled trial than to choose by what
happens to be readily available, hunch, or what a drug company
recommends. The most effective argument in favour of randomised
clinical trials is that the alternative, practising in complacent un-
certainty, is worse. All those points are nicely summarised in the
following quotation from Sir George Pickering, made when President
of the Royal Society of Medicine in 1949, in response to the charge
that the clinical trial constituted experimentation on patients:

All therapy is experimentation. Because what in fact we are doing is to


alter one of the conditions, or perhaps more than one, under which our pa-
tient lives. This is the very nature of an experiment., because an experiment
is a controlled observation in which one alters one or more variables a t a
time to try to see what, happens. The difference between haphazard therapy
and a controlled clinical trial is that in haphazard therapy we carry out our
experiments without, design on our patients and therefore our experiments
An Introduction to Clinical Dials 15

are bad experiments from which it is impossible to learn. The controlled


clinical trial merely means introducing the ordinary accepted criteria of a
good scientific experiment.

Further convincing empirical arguments in favour of the double-


blind controlled clinical trial are provided by the work of Chalmers
et al. (1977) and Sacks et al. (1983) who provide evidence that
nonrandomised studies yield larger estimates of treatment effects
than studies using random allocation (see Table 1.3), estimates that
are very likely biased; and Schulz et al. (1995), who demonstrate
that trials in which concealment of treatment allocation was either

Table 1.3. Results from Randomised and Historical Control


Trials in Six Areas.
Randomised Trials Historical Control Trials
New treat. New treat. New treat. New treat.
Therapy effective ineffective effective ineffective

Coronary artery 1 7 16 5
surgery.
Anticoagulants for 1 9 5 1
acute myocardial
infarction.
Surgery for 0 8 4 1
oesophaeal varices.
Flurouracil (5-FU) 0 5 2 0
for colon cancer.
BCG 2 2 4 0
immunot her apy
for melanoma.
Diethylstilbesterol 0 3 5 0
for habitual
abortion.

(Taken from Sacks e t al., 1983.)


16 Design and Analysis of Clinical Trials

inadequate or unclear (i.e., were not double-blind), also yielded larger


(biased) estimates of treatment effects.
There are a number of other ethical issues in clinical trials which
relate directly to one or the other of the statistical aspects of design
and analysis; an exampIe is determining the appropriate sample size
by means of a power analysis - using too small or too large a sample
would be unethical, a point that will be taken up in more detail in
the next chapter.

1.5. CLINICAL TRIAL PROTOCOLS


All clinical trials begin wit,h a protocol which serves as a guide for
the conduct of the trial. The protocol must describe in a clear and
unambiguous manner how the trial is to be performed so that all the
investigators are familiar with the procedures to be used. The proto-
col must summarise published work on the study topic and use the
results from such work to justify the need for the trial. If drugs are
involved, then pertinent pharmacological and toxicity data should
be included. The purpose of the trial and its current importance
need to be described in clear and concise terms; hypotheses that the
trial is designed to test need to be clearly specified and the population
of patients to be entered into the trial fully described. The proto-
col must. specify the treatments to be used; in particular, for drug
studies, the dose to be administered, the dosing regimen, and the
duration of dosing all need to be listed. Details of the randomisation
scheme to be adopted must be made explicit in the protocol along
with other aspects of design such as control groups, blinding, sample
size determination and the number of interim analyses planned (if
any). Although it is important that investigators adhere to the pro-
tocol, mechanisms need to be in place for making changes if the need
arises. If changes are made, then they must be well documented.

1.6. SUMMARY
The controlled clinical trial has become one of the most important
tools in medical research and investigators planning to undertake
An Introduction to Clinical Trials 17

such a trial have no shortage of excellent books to which to turn for


advice and information. But unlike the many other books dealing
with clinical trials, this text is primarily concerned with the statisti-
cal issues of certain aspects of their design (Chapters 2 and 3) and,
in particular, their analysis (Chapters 4 to lo), rather than their
day-to-day organisation. This restriction will enable us to give fuller
accounts of some recently developed methods that may be particu-
larly useful for the type of data often generated from clinical trials.
Some details of the software available that implements the methods
described will be given in the Appendix.
CHAPTER2

Treatment Allocation,
the Size of Trials
and Reporting Results

2.1. INTRODUCTION
The design and organisation of a clinical trial generally involves a
considerable number of issues. These range from whether it is ap-
propriate to mount the trial at all, to selection of an appropriate
outcome measure. Some of these issues will be of more concern to
statisticians than others. Three of these: 1) allocating subjects to
treatment groups; 2) deciding the size of the trial, i.e., how many
subjects should be recruited; and 3) how results should be reported,
will be discussed in this chapter.

2.2. TREATMENT ASSIGNMENT


METHODS
One of the most important aspects of the design of a clinical trial
is the question of how patients should be allocated to the various
treatments under investigation. As Silverman (1985) put it:

How is the impossible decision made to choose between the accepted


standard treatment and the proposed improved approach when a fellow
human being must be assigned to one of the two (or more) treatments

18
Treatment Allocation, the Size of Trials and Reporting Results 19

under test? Despite the most extensive pre-clinical studies, the first human
allocation of a powerful treatment is largely a blind gamble and it is perhaps
not surprising that so much has been written on the most appropriate
fashion to allocate treatments in a trial.

Most of the early clinical experiments involved arbitary, nonsys-


tematic schemes for assigning patients t o treatments (see, for ex-
ample, the description of Lind’s experiment in Chapter 1). The
concept of randomisation as a device for treatment assignment was
introduced by Fisher in the 1920s (all-be-it in the context of agri-
cultural experimentation). Randomised trials have been in existence
since the 1940s but only in the last 10 or 20 years have they gained
widespread acceptance. Possible ethical objections t o randomisation
were mentioned in Chapter 1, and many alternative methods of al-
location have been suggested, the defficiencies of most of which are
well documented in, for example, Pocock (1983). T h e main com-
petitor t o randomisation is the use of historical controls; all suitable
patients receive the new treatment and their outcomes are compared
with those from the records of patients previously given the standard
treatment. Although there are considerable problems with the use of
such controls (see Table 2.1), it has now become more widely recog-
nised that they do have a role to play (see, for example, Simon, 1982),

Table 2.1. Problems with Historical Control Trials.


~~ ~~~~~ ~~~~ ~

0 Past observations are unlikely to relate to a precisely similar group


of patients
0 The quality of information extracted from the historical controls is
likely to be different (probably inferior), since such patients were not
intially intended to be part of the trial
0 Patients given a new, and as yet, unproven treatment, are likely to
be far more closely monitored and receive more intensive ancillary
care than historical controls receiving the orthodox treatment in a
routine manner
0 For the reasons listed, studies with historical controls are likely to
exaggerate the value of a new treatment (see Table 1.3).
20 Design and Analysis of Clinical Trials

particularly in providing the data on innovative treatments which


might support further investigation in a randomised trial. Some
advantages of historical control studies, particularly in cancer trials,
listed by Gehan (1984) are:

0 historical control studies can cost less than half as much as


randomised studies with comparable sample sizes,
0 there may be many more control patients available historically
than concurrently,
0 a clinician who believes (however weakly) that the experimen-
tal therapy is better than the control faces no ethical dilemma
in admitting patients for treatment,
0 patients are more apt to enlist for a study in which treatment
assignment is not randomised.

But in this text our major concern is with properly randomised


trials. The qualifier is needed here, since the acceptance of the prin-
ciple of randomisation remains only a starting point in the execu-
tion of a trial. If the randomisation is not performed correctly in
practice, then there is every danger that the trial could suffer from
the same biases that are generally suspected in a trial not involving
randomisat ion.
Randomisation in a two-group study might appear to be simply
a matter of repeatedly tossing a coin in order to decide which of
the two treatments each patient should receive, and indeed, in some
circumstances this may be all that is needed. There are, however,
other more complex randomisation schemes designed to achieve var-
ious objectives. We begin, though, with a few comments about the
simple ‘coin tossing’ type of randomisation process. (Although we
shall concentrate largely on trials in which the unit of randomisation
is the individual patient, it is important to note that there are tri-
als in which, for example, complete families are randomised to the
various treatments t o be compared. Such trials may need special
methods of analysis as we shall mention later.)
Treatment Allocation, the Size of Trials and Reporting Results 21

2.2.1. Simple Randomisation


For a randomised trial with two treatments, A and B, the basic con-
cept of tossing a coin over and over again and allocating a patient to
A if a head appears and B if the coin shows tails, is quite reasonable,
but is rarely if ever used in practice. Instead, a randomisation list is
constructed using a published table of random numbers or, more usu-
ally, a computer-based recognised pseudo-random number generator.
The entries in this list can then be used one at a time as patients
are recruited to the trial. (In a multicentre trial, each centre should
have its own randomisation schedule so as to avoid treatment-centre
confounding.)
The advantage of such a simple method is that each treatment as-
signment is completely unpredictable and in the long run the number
of patients allocated to each treatment is unlikely to differ greatly.
In the long run, however, implies a greater number of patients than
are recruited to many clinical trials and it is of some interest to

Table 2.2. Possible Imbalance in Simple Randomisat ion


with Two Treatments.
Shows the difference in treatment numbers (or more
extreme) liable to occur with probability at least 0.05 or
at least 0.01 for various trial size.
Total Number Difference in Numbers
of Patients Probability 2 0.05 Probability 2 0.01

10 2:8 1 :9
20 6:14 4:16
50 18:32 16:34
100 40:60 37:63
200 86:114 82:118
500 228:272 221:279
1000 469:531 459:541
(Taken with permission from Pocock, 1983.)
22 Design and Analysis of Clinical Trials

consider the chance of possible imbalance of patient numbers in the


two groups. Table 2.2 (taken from Pocock, 1983), illustrates the
differences in treatment numbers that may occur with probability
greater than 0.05 and 0.01. For a trial with 20 patients, for ex-
ample, the chance of four being allocated to one treatment and 16
to the other is greater than 0.01. Although this chance might be
regarded as fairly small, the resulting imbalance would be of grave
concern to most investigators, typically resulting in a study of much
lower power than that expected from an even allocation of subjects
(although see Section 2.2.5). Consequently, it is often desirable to
restrict randomisation to ensure similar treatment numbers through-
out the trial. (Although see later for situations when unequal group
sizes may be a sensible feature of the design.) Several methods are
available for achieving balanced group sizes of which the most com-
monly used is blocked randomisation.

2.2.2. Blocked Randomisation


This method introduced by Hill (1951), and also known as permuted
block randomisation, guarantees that at no time during randomisa-
tion will the imbalance be large and that at certain points the number
of subjects in each group will be equal. The essential feature of this
approach is that blocks of a particular number of patients are con-
sidered and a different random ordering of treatments assigned in
each block; the process is repeated for consecutive blocks of patients
until all have been randomised. For example, with two treatments
(A and B), the investigator may want t o ensure that after every sixth
randomised subject, the number of subjects in each treatment group
is equal. Then a block of size 6 would be used and the process would
randomise the order in which three As and three Bs are assigned
for every consecutive group of six subjects entering the trial. There
are 20 possible sequences of 3As and 3Bs, and one of these is cho-
sen at random, and the six subjects are assigned accordingly. The
process is repeated as many times as possible. When six patients are
Treatment Allocation, the Size of Trials and Reporting Results 23

enrolled, the numerical balance between treatment A and treatment


B is equal and the equality is maintained with the enrollment of the
12th, 18th, etc., patient.
Freidman, Furberg and DeMets (1985) suggest an alternative
method of blocked randomisation in which random numbers between
0 and 1 are generated for each of the assignments within a block, and
the assignment order then determined by the ranking of these num-
bers. For example, with a block of size six in the two treatment
situation, we might have:

Assignment Random Number


A 0.112
A 0.675
A 0.321
B 0.018
B 0.991
B 0.423

This leads to the assignment order BAABAB.


In trials that are not double-blind, one potential problem with
blocked randomisation is that at the end of each block an investigator
who keeps track of the previous assignments could predict what the
next treatment would be. This could permit a bias to be introduced.
The smaller the block size, the greater is the risk of the randomisa-
tion becoming predictable. For this reason, repeated blocks of size
two should not be used. A required means of reducing the prob-
lem is by varying the size of consecutive sets. A random order of
block size makes it very difficult to determine the next assignment in
a series.
The great advantage of blocking is that balance between the
number of subjects is guaranteed during the course of the randomi-
sation. The number in each group will never differ by more than
24 Design and Analysis of Clinical Trials

b / 2 , where b is the size of the block. This can be important for two
reasons. First if enrollment in a trial takes place slowly over a period
of months or even years, the type of patient recruited for the study
may change during the entry period (t,emporal changes in severity of
illness. for example, are not uncommon), and blocking will produce
more comparable groups. A second advantage of blocking is that
if the trial should be terminated before enrollment is completed be-
cause of the results of some form of interim analysis (see Chapter 3),
balance will exist in terms of the number of subjects randomised t o
each group.

2.2.3. Stratified Randomisation


One of the objectives in randomising patients to treament groups
is to achieve between group comparability on certain relevant pa-
tient characteristics usually known as prognostic factors. Measured
prior to randomisation, these are factors which it is thought likely
will correlate with subsequent patient response or outcome. As rnen-
tioned in Chapter 1, randomisation tends to produce groups which
are, on average, similar in their entry characteristics, both known
and unknown. The larger a trial is, the less chance there will be
of any serious non-comparability of treatments groups, but for a
small study there is no guarantee that all baseline characteristics
will be similar in the two groups. If prognostic factors are not evenly
distributed between treatment groups, it may give the investigator
cause for concern. although methods of statistical analysis such as
analysis of covariance exist which allow for such lack of comparablity
(see later chapters). Stratified randomzsation is a procedure which
helps to achieve comparability between the study groups for a chosen
set of prognostic factors. According to Pocock (1983), the method
is rather like an insurance policy in that its primary aim is to guard
against the unlikely event of the treatment groups ending up with
some major difference in patient characteristics.
Treatment Allocation, the Size of Trials and Reporting Results 25

The first issue to be considered when stratified randomisation


is contemplated, is which prognostic factors should be considered.
Experience of earlier trials may be useful here. For example, Stan-
ley (1980) carried out an extensive study of prognostic factors for
survival in patients with inoperable lung cancer based on 50 such
factors recorded for over 5000 patients in seven trials. He showed
that performance status, a simple assessment of the patient’s ability
to get around, was the best indicator of survival. Weight loss in the
last six months and extent of disease also affected survival. These
three factors would, consequently, be those to account for in any
future trial.
When several prognostic factors are to be considered, a stratum
for randomisation is formed by selecting one subgroup from each of
them (continuous variables such as age are divided into groups of
some convenient range). Since the total number of strata is, there-
fore, the product of the number of subgroups in each factor, the
number of strata increases rapidly as factors are added and the lev-
els within factors are refined. Consequently, only the most important
variables should be chosen and the number kept to a minimum.
Within each stratum, the randomisation process itself could be
simple randomisation, but in practice most clinical trials will use
some blocked randomisation approach. As an example, suppose that
an investigator wishes to stratify on age and sex, and to use a block
size of 4. First, age is divided into a number of categories, say 40-
49, 50-59 and 60-69. The design thus has 3 x 2 strata, and the
randomisation might be:

Strata Age Sex Group Assignment


1 40-49 Male ABBA BABA . . .
2 40-49 Female
3 50-59 Male
4 50-59 Female
5 60-69 Male
6 60-69 Female
Design and Analysis of Clinical Trials

Patients between 40-49 years-old and male, would be assigned


to treatment groups A and B in the sequences ABBA BABA . . . .
Similarly random sequences would appear in the other strata.
Although the main argument for stratified randomisation is that
of making the treatment groups comparable with respect to spe-
cific prognostic factors, it may also lead to increased power (see Sec-
tion 2.3) if the stratification is taken into account in the analysis, by
reducing variability in group comparisons. Such reduction allows a
study of a given size to detect smaller group differences in outcome
measures, or to detect a specified difference with fewer subjects.
Stratified randomisation is of most relevance in small trials, but
even here it may not be profitable if there is uncertainty over the
importance or reliability of prognostic factors, or if the trial has a
limited organisation that might not cope well with complex randomi-
sation procedures. In many cases it may be more useful to employ a
stratified analysis or analysis of covariance, to adjust for prognostic
factors when treament differences are assessed (see later chapters).

2.2.4. Minimisation Method


A further approach to achieving balance between treatment groups
on selected prognostic factors is to use an adaptive randomisation
procedure in which the chance of allocating a new patient to a par-
ticular treatment is adjusted according to any existing imbalances
in the baseline characteristics of the groups. For example, if sex is
a prognostic factor and one treatment group has more women than
men, the allocation scheme is such that the next few male patients
are more likely to be randomised into the group that currently has
fewer men. This method is often referred to as minimisation, because
imbalances in the distribution of prognostic factors are minimised
according to some criterion.
In general, the method is applied in situations involving several
prognostic factors and patient allocation is then based on the aim of
balancing the marginal treatment totals for each level of each factor.
Treatment Allocation, the Size of Trials and Reporting Results 27

Table 2.3. Treatment Assignments by Four Prognostic


Factors for 80 Patients in a Breast Cancer Trial.
~ ~~

Factor Level A B
Performance status Ambulatory 30 31
Non-ambulatory 10 9
Age < 50 18 17
2 50 22 23
Disease-free interval < 2 years 31 32
2 2 years 9 8
Dominant metastatic lesion Visceral 19 21
Osseous 8 7
Soft tissue 13 12
(Taken with permission from Pocock, 1983.)

The following example, taken from Pocock (1983), illustrates the


procedure:
Table 2.3 shows the number of patients on each of two treatments,
A and €3, according to each of four prognostic factors. Suppose the
next patient is ambulatory, age < 50, has disease-free interval 2 2
years and visceral metastasis. Then for each treatment, the number
of patients in the corresponding four rows of the table are added:

sum for A = 30 + 18 + 9 + 19 = 76
sum for B = 31 + 17 + 8 + 21 = 77

Minimisation requires the patient be given the breatment with t h e


smallest marginal total, in this case treatment A. If the sums for A
and B were equal, then simple randomisation would be used to assign
the treatment.

2.2.5. Unequal Randomisation


Equal-sized treatment groups provide the most efficient means of
treatment comparison for any type of outcome measure, and the
28 Design and Analysis of Clinical Trials

methods of randomisation described in the previous sections are


aimed at achieving what is, in most circumstances, this desirable
feature of a trial. (In addition, ideally, each centre in a multi-
centre trial should contribute the same number of patients.) But
despite the obvious statistical advantages of groups of equal size,
there may be other considerations which might require more patients
in one group than another. If, for example, the trial is comparing
a new treatment against a standard, the investigator might be far
more interested in obtaining information about the general charac-
teristics of the new treatment, for example, variation in response with
dose, than for the old, where such characteristics are likely to be well
known. (This is often the situation in early Phase I1 trials, for which
there exist historical data on the standard treatment.) Such infor-
mation might be best gained by use of an unbalanced design which
involves allocating a larger number of patients to the new treatment

1.0,

Percentage of patients on the ow treatment

Fig. 2.1. Reduction in power of a trial as the proportion on the new treat-
ment is increased (taken with permission from Pocock, 1983).
Freatment Allocation, the Site of Trials and Reporting Resalts 29

than to the old. Pocock (1996) gives an example of this approach in


a trial for the treatment of advanced breast cancer.
The loss of statistical efficiency in unequal randomisation is con-
sidered by Pocock (1983), who shows that if the overall size of a trial
is kept constant., its power decreases relatively slowly in the move
away from equalsized groups. Figure 2.1 (taken from Pocock, 1983)
demonstrates, for example, that power decreases from only 0.95 t o
0.925 if 67% of patients are allocated to the new treatment.
The more complex the design, the more difficult it becomes to
maintain the integrity of the randomisation process; it becomes
vulnerable t o both accidental and deliberate misallocation. Con-
sequently, there is much to be said for removing the randomisation
decision away from the point of clinical contact, for example, through
the use of specialised 24-hour telephone randomisation schemes.

2.3. THE SIZE OF A CLINICAL TRIAL


According to Simon (1991)

An effective clinical trial must ask a n important question and provide


a reliable answer. A major determinant, of the reliability of the answer is
t,he sample size of the trial. Trials of inadequate size may cause contradic-
tory and erroneous results and thereby lead to an inappropriate treatment
of patients. They also divert limited resources from useful applications
and cheat the patients who participated in what bhey thought was impor-
tant dinical research. Sample size planning is, therefore, a key component
of clinical trial methodology.

But, in the recent past at least, such advice has often not been
heeded and the literature is dotted with accounts of inconsequential
trials. Freiman et al. (1978), for example, reviewed 71 ‘negative’
randomised clinical trials, i.e., trials in which the observed differences
between the proposed and control treatments were not large enough
to satisfy a specified ‘significance’level (the risk of a type I error), and
the results were declared to be ‘not statistically significant’. Analysis
30 Design and Analysis of Clinical Trials

of these clinical studies indicated that the investigators often worked


with numbers of enrolled patients too small to offer a reasonable
chance of avoiding the opposing mistake, a type I1 error. Fifty of
the trials had a greater than 10% risk of missing a substantial dif-
ference (true treatment difference of 50%) in the treatment outcome.
The reviewers warned that many treatments labelled as ‘no differ-
ent from control’ had not received a cribical test, because the trials
had insufficient power t o do the job int.ended. Such trials are, in a
very real sense, unethical in that they require patients to accept the
risks of treatment, however small, without any chance of benefit to
them or future patients. Small-scale preliminary investigations may
be justified when part of a larger plan, but not as an end in their
own right (although see later comments).
A specific example of this problem is given by Andersen (1990),
who reports a study from the New England Journal of Medicine,
in which 52 patients with severe cirrhosis and variceal haemorrhage
requiring six or more pints of blood were randomly assigned either
to sclerotherapy or potocaval shunt. There was no difference in short
term survival, with 13 patients in the sclerotherapy group discharged
alive, as compared with ten patients in the shunt group. The authors
commented as €oHows:

We failed to demonstrate any significant difference in long-term survival


. . . endoscopic sclerotherapy is at least as good as, and may well be better
than, definitive early surgical shunting.

The absence of any significant difference made the investigator


conclude that one treatment is at least as effective as the other.
They failed, however, to consider the possibility of an error of the
second kind. A trial with only 26 patients in each group has only a
50-50 chance of a significant result if the true survival rate on one
treatment is 25% and on the other twice as much. Even though
more patients were discharged alive after sclerothera.py than after
Treatment Allocation, the Size of Trials and Reporting Results 31

portocaval shunt in this study, there remains the distinct possibility


that the operation might eventually turn out to be superior.
The deficiency in patient numbers in many clinical trials is, ac-
cording to Pocock (1996) ‘a general phenomena whose full implica-
tions for restricting therapeutic progress are not widely appreciated’.
In the same article Pocock continues:

The fact is that trials with truly modest treatment effects will achieve
statistical significance only if random variation conveniently exaggerated
these effects. The chances of publication and reader interest are much
greater if the results of the trial are statistically significant. Hence the cur-
rent obsession with significance testing combined with the inadequate size
of many trials means that publications on clinical trials for many treat-
ments are likely to be biased towards as exaggeration of therapeutic effect,
even if the trials are unbiased in all other respects.

The combination of high type I1 error rate, publication bias and


the fact that most true treatment advantages are likely to be mod-
est or nonexistent, almost certainly results in a high proportion of
false positives in the medical literature, a point also made in Peto
et al. (1976) and Zelen (1983). Certainly the case against trials with
inadequate numbers of subjects appears strong but perhaps with the
growing use of meta-analysis, a topic to be discussed in Chapter 10,
not as strong as is implied by the previous comments.
So how, in designing a trial, is its size arrived at? Although prac-
tical and ethical issues need to be considered, most determinations of
sample size for a clinical trial are performed, intially at least, in the
more statistical context of the hypothesis testing framework of Ney-
man and Pearson. A null hypothesis is tested at significance level Q
and the sample size is determined to provide power 1 - /3 for reject-
ing the null when a specified alternative hypothesis is true, where
,f3 is the risk of a type I1 error, i.e., accepting the null hypothesis
when it is false. In simple situations the power is a function of three
factors:
32 Design and Analysis of Clinical Trials

0 the significance level adopted,


0 the reliability and variability of the sample data,
0 the size of the treatment effect.
The required sample size will be larger the higher the level of
significance chosen, the lower the reliability and the smaller the treat-
ment difference hypothesised under the alternative hypothesis. In
many trials, the power calculation will also need to include consid-
eration of other factors such as the follow-up time versus number of
cases and the number of measurement points.
In this paradigm, the general approach is for the investigator to
specify the size of the treatment difference considered clinically rel-
evant (i.e., important to detect) and with what degree of certainty,
i.e., with what power, it should be detected. Given such information,
the calculation of the corresponding sample size is often relatively
straightforward, although the details will depend on the type of re-
sponse variable and the type of test involved. The last decade has
produced a large volume of methodology useful in planning the size of
randomised clinical trials with a variety of different types of outcome
measures - some examples are t o be found in Lee (1983), McHugh
and Lee (1984), Schoenfield (1983), Sieh (1987), Wittes and Wallen-
stein (1987) and Spiegelhalter et al. (1994). In many cases, tables
are available which enable the required sample size t o be simply read
off. Increasingly, these are being replaced by computer software for
determining sample size for many standard and non-standard designs
and outcome measures (see the Appendix).
An obvious danger with such a procedure is that investigators
(and, in some cases, their statisticians) may occasionally be led to
specify an alternative hypothesis that is unrealistically extreme so
that the required sample size looks feasible in terms of possible press-
ing temporal and financial constraints. Such a possibility may be
what led Senn (1997) to describe power calculations as ‘a guess mas-
querading as mathematics’ and Pocock (1996) to comment that they
are ‘a game that can produce any number you wish with manipu-
lative juggling of the parameter values’. Statisticians advising on
Treatment Allocation, the Size of Trials and Reporting Results 33

clinical trials need to be active in estimating the degree of difference


that can be realistically expected for a clinical trial based on previous
studies of a particular disease or, when such information is lacking,
perhaps based on subjective opinions of investigators and physicians
not involved in the proposed trial.

2.4. REPORTING RESULTS

The first part of a trial report should be descriptive and summarise


the patient pool, the study protocol and the characteristics of those
entered. Appropriate graphical material should be included here.
Following this will be the results of the various statistical analyses
performed. In many cases the analyses will involve the application
of one or other significance test and what is often tabulated is the p
value associated with the test. Unfortunately and despite the many
caveats in the literature, the accept/reject philosophy of significance
testing remains dominant in the minds of many non-statisticians,
who appear determined to continue to experience joy on finding a
pvalue of 0.049 and despair on finding one of only 0.051. A decade
ago, Gardner and Altman (1986) made the point that the excessive
use of hypothesis testing at the expense of other ways of assessing re-
sults had reached such a degree that levels of significance were often
quoted alone in the main text and abstracts of papers, with no men-
tion of actual concentrations, proportions, etc., or their differences.
The implications of hypothesis testing - that there can always be a
simple ‘yes’ or ‘no’ answer as the fundamental result from a medical
study - is clearly false, and used in this way hypothesis testing is
of limited value.
So should statisticians be encouraging the abandonment of the
pvalue altogether? Many statisticians might be tempted to answer
‘yes’, but a more sensible response is perhaps a resounding ‘maybe’.
Such values should rarely be used in a purely confirmatory way, but
in an exploratory fashion they can give some informal guidance on
the possible existence of an interesting effect, even when the required
34 Design and Analysis of Clinical Dials

assumptions of whatever test is being used are known to be only


partially valid. It is often possible to assess whether a p-value is
likely to be an under- or overestimate, and whether the result is
clear one way or the other.
Fortunately, the use of significance testing appears to have be-
come less obsessive and dogmatic during the last few years, with
greater emphasis on statistical estimation and confidence intervals.
The latter, which can be considered to be the set of true but unknown
differences that are statistically compatible with the observed differ-
ence, can be found relatively simply for many quantities of interest
(see Gardner and Altman, 1986), and although the underlying logic
of interval estimates is essentially similar to that of significance tests,
they do not carry with them the pseudo scientific decision making
language of such tests. Instead they give a plausible range of values
for the unknown parameter, with, for example, inadequate sample
size being signalled by the sheer width of the interval. As Oakes
(1986) rightly comments:

The significance test relates to what a population parameter is not: the


confidence interval gives a plausible range for what the parameter is.

Since clinical trials are generally designed to provide global


treatment comparisons, which may not be suited to the needs of in-
dividual patients, a question which frequently arises when reporting
results, whether as p-values or confidence intervals, is how to iden-
tify particular subgroups of patients who responded well (or badly)
to a new treatment? Answering such a question is relatively easy -
such subgroup analysis can be carried out using standard statistical
techniques such as analysis of variance or the like. If, for example,
subgroups were formed on the basis of sex (male, female) and age
(young, old), then with a two treatment trial, a 2 x 2 x 2 ANOVA on
the response variable of interest could be performed and all possible
interactions and main effects assessed in the usual way. But many
statisticians would recommend that such analyses are better avoided
Treatment Allocation, the Size of Trials and Reporting Results 35

altogether, or if undertaken, interpreted extremely cautiously in the


spirit of ‘exploration’ rather than anything more formal. Their rea-
sons are not difficult to identify:

0 trials can rarely provide sufficient power to detect such sub-


group effects,
0 there are often many possible prognostic factors from which
to form subgroups, so that the analysis may degenerate into
‘data dredging’,
0 the temptation to over interpret an apparent subgroup finding
is likely to be difficult to resist (Yusuf et al., 1991).

Chapter 10 (Fig. 10.2 and Table 10.1) provides a checklist for


what is currently expected that studies should report.

2.5. SUMMARY

In this chapter three aspects of the design of clinical trials has been
considered: the allocation of patients to treatments, the size of the
trial, and how to report results. The well documented chaos that can
result from non-randomised trials has led to a general acceptance of
randomisation as an essential component of the vast majority of tri-
als. Several methods of randomisation have been considered in this
chapter, but there are others that have not been mentioned, in partic-
ular cluster randomisation in which natural groupings of individuals,
for example, general practices, become the units of randomisation,
and play the winner rules, in which the randomisation is biased to
allow a higher proportion of future allocations to the treatment with
(currently) better observed outcome. The former are described in
Donner and Klar (1994) and are increasingly important in, for ex-
ample, country intervention trials. The latter presumably arise from
the desire amongst some clinicians for more ‘ethically acceptable’
allocation procedures, but there have been very few clinical trials
that have actually used such data-dependent procedures. Pocock
36 Design and Analysis of Clinical Trials

(1996) points t o both the impracticality of implementation (early re-


sults usually arrive too late) and the suspicions of bias in allocations
(the result, for example, of secular trends in the type of patient
recruited) as reasons that such approaches continue to be largely re-
garded as ‘statistical curiosities’ rather than serious contenders for
clinical applications.
Statisticians have been very effective in developing methodology
for determining sample size in randomised clinical trials. Neverthe-
less, there is continuing evidence that many reported trials are too
small, leading to a high type I1 error rate as well as a low proportion
of true positive to false positive findings. Meta-analysis (see Chap-
ter 10) may help with this problem, but it is unlikely to provide a
completely satisfactory answer in all cases.
Ten years ago the results of most clinical trials were given in
terms of p-values. The situation has now changed for the better,
with many medical journals rightly demanding confidence intervals.
The problem with significance testing would not be so bad if only
a single test (or a very small number) was carried out per trial.
Most trials, however, generate large amounts of data and dozens of
significance tests from the results of using such procedures as i n t e r i m
analyses and employing multiple endpoints, topics to be discussed in
the next two chapters.
CHAPTER3

Monitoring Trial
Progress: Outcome
Measures, Compliance,
Dropouts and
Interim Analyses

3.1. INTRODUCTION

The basic elements of the plan for any trial will be set long before the
first patient is enrolled, and will, eventually, be translated into the
study protocol. Here the primary objectives of the trial will be de-
tailed in terms of type of patients to be studied, class of treatments
to be evaluat,ed and primary outcome measures. In addition. the
document will specify the number of patients to be recruited (usu-
ally from the results of a sample size calculation), required length
of patient follow-up, pat.ient entry and exclusion crit.eria, method
of randomisation, details of pre-randomisation procedures and other
general organisational structures.
Putting the study plan into execution begins with patient recruit-
ment, a period which is crucial in the life of a trial, although the
details need not delay us here, since the various possible problems
and pitfalls are well documented in, for example, Meinert (1986).

37
38 Design and Analysis of Clinical Trials

Once a trial has started, various aspects of a patient’s progress need


to be assessed. Investigators are, for example, under a strict obliga-
tion to report unexpected adverse events as they occur since these
may necessitate withdrawing of a patient from the trial.
The problem of patient compliance must also be addressed. The
optimal study from a compliance point of view is one in which the
investigator has total control over the patient, the administration
of the intervention regimen and follow-up. But in practice, there is
very likely to be less than 100% compliance with the intervention
and clearly the results of a trial can be affected by noncompliance; it
can, for example, lead to an underreporting of possible therapeutic
as well as toxic effects. This has the potential to undermine even
a properly designed trial and consequently, monitoring compliance
is generally critical in a clinical trial, a point taken up in detail in
Section 3.3.
A patient’s performance during a clinical trial is characterised
by measurements on one or more outcome variables. These measure-
ments need to be made in as objective, accurate and consistent a
manner as possible and in a way that should be precisely defined in
the study protocol, issues that are discussed more fully in Section 3.2.
In most clinical trials, patients are entered one at a time so that
their responses to treatment are also observed sequentially. The ac-
cumulating data in a trial needs to be monitored for a variety of
reasons. In addition to checking for compliance and noting possi-
ble adverse side effects, monitoring is also needed to assess whether
early termination of the trial might be necessary. Early termination
might be called for if there was an indication that the intervention
was harmful to patients. Alternatively, if the data indicate a clear
benefit from the intervention, the trial may need to be stopped early
because to continue to use the control treatment would be unethi-
cal. The handling of treatment comparisons while a trial is still in
progress poses some difficult statistical problems which are taken up
in Section 3.4.
Monitoring Trial Progress 39

3.2. OUTCOME MEASURES


The outcome measure(s) used for treatment comparisons may be a
clinical event, for example, death or recurrence of a disease, or a
measurement of some other characteristics of interest, for example,
blood pressure, serum lipid level or breathing difficulties. Such ob-
servations and measurements are the raw material of the trial and
they clearly need to be objective, precise and reproducible for reasons
nicely summarised by the following quotation from Fleiss (1986):

The most elegant design of a clinical study will not overcome the dam-
age caused by unreliable or imprecise measurement. The requirement that
one’s data be of high quality is at least as important a component of a
proper study design as the requirement for randomisation, double blinding,
controlling where necessary for prognostic factors and so on. Larger sample
sizes than otherwise necessary, biased estimates, and even biased samples
are some of the untoward consequences of unreliable measurements that
can be demonstrated.

So no trial is better than the quality of its data and quality


control begins with clear definitions of response variables. The de-
cision about which, when and how measurements are to be made
needs to be taken before the trial commences. The alternative is
potential chaos. Attention clearly needs to be given to training clin-
icians and others on the measuring instruments to be used; this is
particularly important in multi-centre trials. Results from studies
based on poorly standardised procedures that use ambiguous defini-
tions or conducted by insufficiently trained staff, can lead to both
loss of power and bias in the estimate of treatment effect. Most
properly conducted trials will, in fact, have well developed systems
in place for data quality control and auditing. The purpose of such
a system is to provide reasonable assurance to the organisers of the
trial as well as to the ‘consumers’ of the results that the data on
which the conclusions are based are reliable. Some practical issues
in assuring such quality control of the data generated in clinical trials
are discussed in Knatterud et al. (1998).
40 Design and Analysis of Clinical Dials

The measurements made in many trials will involve rating


scales of one type or another, for example, quality-of-life assessments.
When there has been little experience of using such scales or where
they have only been recently developed, it may be important to in-
vestigate their reliability, i.e. , the extent to which measurements of
the same subject made by different observers agree. In most trials,
particularly multi-centre trials, the problem of observer variation will
need to be confronted. A formal assessment of the reliability of the
instrument to be used may even be necessary. Readers are referred
to the comprehensive accounts of assessing reliability in both Fleiss
(1986) and Dunn (1989) for details.
In some trials the outcome measure of substantive interest may
not be measured for practical and/or ethical reasons. Instead a sur-
rogate variable is selected on which to investigate treatment differ-
ences. If, for example, we measure blood pressure rather than say
the potential problems of high blood pressure, we are using a surro-
gate measure. Other examples include bone mineral density in the
treatment of osteoporosis rather than the rate of bone fracture, and
CD4 counts rather than deaths in the treatment of AIDS.
The aim in using such surrogate measures is to assess the treat-
ment effect with less trouble and perhaps greater efficiency than by
using the preferred endpoint. But the dangers of not using the true
endpoint cannot be dismissed lightly. Senn (1997) points out that
the demonstration of a high correlation between the proposed surro-
gate measure and the measure of substantive interest does not ensure
that using the former will be adequate. He uses the treatment of
oestoporosis measured by bone mineral density (BMD) as an exam-
ple. Loss of BMD leads to a weakening of bones and an increased
risk of fracture. If, however, a treatment increases density but at
the expense of adversely affecting the construction of the bones it
may actually have a harmful effect on the fracture rate. The im-
plication is that the adequacy of a surrogate measure may depend
on the treatment, a possibility recognised by the USA drugs regula-
tory body, the Food and Drug Administration (FDA) guidelines for
Monitoring Trial Progress 41

oesteoporosis studies, these requiring the measure of fractures for


bisphophonates but accepting BMD for hormone replacement.

3.3. COMPLIANCE, DROPOUTS AND


INTENTION-TO-TREAT
There could be no worse experimental animals on earth than human beings;
they complain, they go on vacations, they take things they are not supposed
to take, they lead incredibly complicated lives, and, sometimes, they do
not take their medicine. (Efron, 1998.)

Compliance means following both the intervention regimen and


trial procedures (for example, clinic visits, laboratory procedures and
filling out forms). A non-complier is a patient who fails to meet the
standards of compliance as established by the investigator. A high
degree of patient compliance is an important aspect of a well-run
trial.
But treatment compliance is rarely an all-or-none phenomena.
The level of compliance achieved may range from low to high, de-
pending on both the patient and the staff. Perfect compliance is
probably impossible to achieve, particularly in drug trials where
the patient maybe required to take the assigned medication at the
same time of day over long periods of time. Lack of compliance can
take a number of forms; the patient can drop out of the trial, take
none of the medication (whilst perhaps pretending to do so), forget
to take the treatment from time to time, or take it at the wrong
time, etc.
Level of compliance will depend on a number of factors, including:

0 The amount of time and inconvenience involved in making


follow-up visits to the clinic,
0 The perceived importance of the procedures performed at each
visit from a health maintenance point of view,
0 The potential health benefits associated with treatment versus
potential risks,
42 Design and Analysis of Clinical Trials

0 The amount of discomfort produced by the study treatments


or procedures performed,
0 The amount of effort required of the patient to maintain the
treatment regime,
0 The number and type of side effects associated with treatment.

In recent times the problems of noncompliance in a clinical trial


have been well illustrated in trials involving HIV/AIDS patients,
where an atmosphere of rapidly alternating hopes and disappoint-
ments has added t o the difficulties of keeping patients on a fixed
long-term treatment schedule.
So what can be done to ensure maximal patient compliance? As-
pects of the study design may help; the shorter the trial, for example,
the more likely subjects are to comply with the intevention regimen.
So a study started and completed in one day would have great advan-
tages over longer trials. And studies in which the subjects are under
close supervision, such as in-patient hospital-based trials, tend to
have fewer problems of noncompliance.
Simplicity of intervention may also affect compliance, with single
dose drug regimens usually being preferable to those requiring multi-
ple doses. The interval between scheduled visits to hospital or clinic
is also a factor to consider. Too long an interval between visits may
lead to a steady fall in patient compliance due to lack of encourage-
ment, while too short an interval may prove a nuisance and reduce
cooperation.
Perhaps the most important factor in maintaining good subject
compliance once a trial has begun is the attitude of the staff running
the trial. Experienced investigators stay in close contact with the
patients early after randomisation to get patients involved and, later,
to keep them interested when their initial enthusiasm may have worn
off. On the other hand, uninterested or discourteous staff will lead
to an uninterested patient population. Meinert (1986) lists a number
of simple factors likely to enhance patient participation and interest;
this list is reproduced here in Table 3.1.
Monitoring Trial Progress 43

Table 3.1. Factors and Approaches that Enhance Patient


Interest and Participation.
Clinic staff who treat patients with courtesy and dignity and who
take an interest in meeting their needs,
Clinic located in pleasant physical surroundings and in a secure en-
vironment,
Convenient access to parking for patients who drive, and t o other
modes of transportation for those who do not,
Payment of parking and travel fees incurred by study patients,
Payment of clinic registration fees and costs for procedures required
in the trial,
Special clinics in which patients are able to avoid the confusion and
turmoil of a regular out-patient clinic,
Scheduled appointments designed to minimise waiting time,
Clinic hours designed for patient convenience,
Written or telephone contacts between clinic visits,
Remembering patients on special occasions, such as Christmas,
birthday anniversaries, etc.,
Establishment of identity with the study through proper indoctri-
nation and explanantion of study procedures during the enrollment
process; through procedures such as the use of special ID cards to
identify the patient as a participant in the study, and by awarding
certificates to recognise their contributions to the trial.
(Taken with permission from Meinert, 1986.)

Monitoring compliance is a crucial p a r t of m a n y clinical trials,


since according t o Freidman, Furberg and DeMets (1985):

. . . the interpretation of study results will be influenced by knowledge


of compliance with the intervention. To the extent that the control group
is not truly a control group and the intervention group is not being treated
as intended, group differences may be diluted, leading possibly to an under-
estimate of the therapeutic effect and an underreporting of adverse effects.

Feinstein (1974) points o u t that differential compliance to two


equally effective regimens can also lead t o possibly erroneous conclu-
sions about the effect of the intervention.
44 Design and Analysis of Clinical Trials

In some studies measuring compliance is relatively easy. For ex-


ample, trials in which one group receives surgery and the other group
does not. Most of the time, however, assessment of compliance is not
so simple and can rarely be established perfectly. In drug trials one
of the most commonly used methods of evaluating subject compli-
ance is pill or capsule count. But the method is far from foolproof.
Even when a subject returns the appropriate number of leftover pills
at a scheduled visit, the question of whether the remaining pills were
used according to the protocol remains largely unanswered. Good
rapport with the subjects will encourage cooperation and lead to a
more accurate pill count, although there is considerable evidence that
shows that the method can be unreliable and potentially misleading
(see, for example, Cramer et al., 1988, and Waterhouse et al., 1993).
Laboratory determinations can also sometimes be used to moni-
tor compliance to medications. Tests done on either blood or urine
can detect the presence of active drugs or metabolites. For exam-
ple, Hjalmarson et al. (1981) checked compliance with metroprobol
therapy after myocardial infarction by using assays of metroprobol
in urine. Several other approaches to monitoring compliance are de-
scribed in Freidman, Furberg and Demets (1985), and Senn (1997)
mentions two recent technical developments which may be useful,
namely:

0 Electronic monitoring ~ pill dispensers with a built-in micro-


chip which will log when the dispenser was opened,
0 Low-dose, slow turnover chemical markers which can be added
to treatment and then detected via blood-sampling.

The claim is often made that in published drug trials more than
90% of patients have been satisfactorily compliant with the protocol-
specified dosing regimen. But Urquhart and DeKlerk (1998) sugest
that these claims, based as they usually are, on count of returned
dosing forms, which patients can easily manipulate, are exaggerated,
and that data from the more reliable methods for measuring compli-
ance mentioned above, contadict them.
Monitoring Dial Progress 45

Noncompliance may lead to the investigator transferring a patient


to the alternative therapy or withdrawing the patient from the study
altogether; often such decisions are taken out of the investigators
hands by the patient simply refusing to participate in the trial any
further and thus becoming a trial dropout. When noncompliance
manifests as dropout from a study, the connection with missing data
is direct (see Section 3.3.2). In other circumstances manifestation
of noncompliance is more complex and some response is observed,
but a question remains about what would have been observed had
compliance been achieved.
Noncompliance, leading either to receiving treatment other than
that provided for by the results of randomisation, or to dropping out
of the trial altogether, has serious implications for the analysis of the
data collected in a clinical trial, implications which will be discussed
briefly here and taken up again in later chapters.

3.3.1. Intent ion-t +Treat


As indicated above, in most randomised clinical trials not all patients
adhere to the therapy to which they were randomly assigned. Instead
they may receive the therapy assigned to another treatment group, or
even a therapy different from any prescribed in the protocol. When
such non-adherence occurs, problems arise with the analysis compar-
ing the treatments under study. There are a number of possibilities
of which the following are the most common:
Intention-to-treat or analysis-as-randomised in which analysis
is based on original treatment assignment rather than treat-
ment actually received,
Adherers-only method, i.e. , analysing only those patients who
adhered to the original treatment assignment,
Treatment-received method, i.e., analysing patients according
to the treatment ultimately received.
The intention-to-treat (ITT) approach requires that any com-
parison of the treatments is based upon comparison of the outcome
46 Design and Analysis of Clinical Trials

results of all patients in the treatment groups to which they were


randomly assigned. This approach is recommended since it main-
tains the benefits of randomisation, whereas the second and third
of the methods above compare groups that have not been randomised
to their respective treatments; consequently, the analyses maybe sub-
ject to unknown biases and for this reason most statisticians and drug
regulatory agencies prefer intention-to-treat. But although it is clear
that analyses based on compliance are inherently biased because non-
compliance does not occur randomly, many clinicians (and even some
statisticians) have criticised analysis that does not reflect the treat-
ment actually received, especially when many patients do not remain
on the initially assigned therapy (see, for example, Feinstein, 1991).
In the face of substantial non-compliance, it is not difficult to under-
stand the intuitive appeal of comparing only those patients in the
original trial that actually complied with the prescribed treatment.
However, in addition to the difficulty of defining compliance in an
objective manner, subjects who comply tend to fare differently and
in a somewhat unpredictable way from those who do not comply.
Thus any observed difference among treatment groups constructed
in this way may be due not to treatment but to factors associated
with compliance.
Dissatisfaction with analysis by original treatment assignment
arises because of its apparent failure to evaluate the ‘true’ effect of the
treatment. According to Dixon and Albert (1995), an intention-to-
treat analysis determines treatment effectiveness where this involves
both compliance on treatment, as well as its biological effect, whereas
an as-treated analysis assesses treatment efficacy. This, however,
appears to simply be ignoring the potential problem of bias in the
latter.
Peduzzi et al. (1993) compare the various methods of analysis
on data from a randomised trial of coronary artery bypass surgery
designed to compare the survival times of patients assigned opti-
mal medical therapy with those assigned coronary bypass surgery.
Monitoring Trial Progress 47

Amongst the 354 patients assigned to medical therapy, the cumula-


tive 14-year crossover rate from medical to surgical therapy was 55%.
In contrast, only 20 of the 332 patients assigned to surgical therapy
refused surgery. Analysis by the as-randomised approach indicated
that the treatment groups were statistically indistinguishable. In
contrast the analyses by adherers-only and by treatment received
indicated an apparent consistent survival advantage with surgical
therapy throughout the entire 14-year follow-up period, although the
advantage begins to diminish with extended follow-up. The authors
demonstrate that the apparent survival advantage arises from the
striking difference in survival between the crossovers to surgery and
the medical adherers.
In the same paper, some simulated data are used to empha-
sise the problems with other than an as-randomised analysis. One
example presented involves simulated data for a hypothetical co-
hort of 350 medical and 350 surgical patients having exponentially
distributed survival times and assuming a 10-year survival rate of
50% in each group. In addition, they generated an independent ex-
ponential time to ‘crossover’ for each of the 350 medical patients as-
suming half the patients crossed over by 10 years. Medical crossovers
were then defined as those patients with time to crossover less than
survival time. Figure 3.1 displays 10-year survival rates by the as-
randomised, adherers-only, and treatment-received methods. The
latter two methods demonstrate a consistent survival advantage in
favour of surgical therapy, when by definition here, there is actually
no difference in survival between the two treatment groups.
According to Efron (1998), ‘Statistics deals with the analysis of
complicated noisy phenomena, never more so than in its applications
to biomedical research, and in this noisy world the intent-to-treat
analysis of a randomised double-blinded clinical trial stands as a
flagpole of certainty amongst the chaos.’ Indeed, according to Goet-
ghebeur and Shapiro (1993), intention-to-treat analysis has achieved
the status of a ‘Buick’ - ‘Best Unbiased Inference with regard to
Causal Knowledge’. Many statisticians would endorse these views
48 Design and Analysis of Clinical Trials

Treatment Group Control Group

i
+

. .
. .' . *-

Fig. 3.1. Compliance dose-response curves for decrease in cholesterol level


in active treatment and control groups in a trial of cholestyramine (taken
with permission from Peduzzi et al., 1993).

and also find themselves largely in agreement with Peduzzi et al.


(1993):

We conclude that the method of analysis should be consistent with the


experimental design of a study. For randomised trials, such consistency
requires the preservation of the random treatment assignment. Because
methods that violate the principles of randomisation are susceptible to bias,
we are against their use.

But despite widespread agreement amongst statisticians that


intention-to-treat analysis remains the most appropriate way to deal
with noncompliance, there is growing interest in how to take com-
pliance information into account without fatally compromising the
conclusions of a randomised clinical trial, particularly now that mea-
surement of compliance can be made more reliable. (Dunn, 1999,
considers the problem of measurement error in assessing compliance.)
There have been several attempts to incorporate compliance data
into the analysis of clinical trials and to devise analytic methods
Monitoring Trial Progress 49

AS - RANDOMIZED ADHERERS - ONLY TREATMENT - RECEIVED

‘\

P - ’OOI2’
-_
? < .OOOt’
5 0.3
MEDICAL IN
S U O C A L IN
-- 3501
3501
---- MEDICAL IN
sunaicAL IN
-- 2111
3501 -
MEDICAL IN
SURGICAL IN
-- 2111
4141
U
0.1

0.0
0 2 4 I b 10 0 2 4 I $ 1 0 0 1 , 8 8 10

YEARS ON STUDY YEARS ON STUDY YEARS ON STUDY

Fig. 3.2. Ten-year survival rates by the as-randomised, adherers-only and


treatment-received approaches for a set of simulated data (taken with per-
mission from Efron and Feldman, 1991).

that adjust for noncompliance. Examples include Efron and Feld-


man (1991), Pocock and Abdalla (1998), and Robins (1998). Efron
and Feldman, for example, discuss a trial concerned with the effec-
tiveness of the drug cholestyramine for lowering cholesterol levels,
in which each patient’s compliance was assessed by the proportion
of the intended dose actually taken. Figure 3.2 shows the relation-
ship between compliance and the decrease in cholesterol level for
both treatment and control groups. A ‘dose-response’ relation is ev-
ident for both groups; better compliance leads to a greater decrease
in cholesterol level, as indicated by the quadratic regression curves.
But the curves shown are compliance dose-response curves; they may
not give an accurate picture of the true dose-response curve because
compliance (and hence dose) has not been assigned in a randomised
fashion by the investigators. Compliance is an uncontrolled covari-
ate, and it may be that better compliers are better patients to be-
gin with. This seems very likely given the nature of the observed
curve in the control group. Efron and Feldman describe how the
true dose-response curve can be recovered from the treatment and
control compliance-response curves.
50 Design and Analysis of Clinical Trials

Pocock and Abdalla (1998), whilst accepting that analysis by


intention-to-treat remains the statistical approach for presenting the
comparative results of different treatment policies within a random-
ised controlled trial, comment on the growing interest in exploring
more complex statistical approaches which incorporate measures of
individual patient compliance with the intended treatment regimens
into supplementary comparative analyses. They offer an example
from their own analysis of a three-arm study of cardiology patients,
testing a beta-blocker and a diuretic versus placebo. As expected, the
diuretic group showed a tendency to increased serum cholesterol lev-
els, but unexpectedly the same effect showed up in the beta-blocker
group. It was discovered, however, that 30% of the beta-blocker
group was also taking diuretics, and the beta-blocker cholesterol ef-
fect disappeared when this fact was incorporated in the analysis.
In the past the inclusion of compliance measurements in the anal-
ysis of clinical trials has been fiercely resisted by many statisticians,
and if not resisted, largely ignored. But in Efron’s view (Efron,
1998), this will change and at some time in the not too distant fu-
ture, it will seem as wrong to run a clinical trial without compliance
measurement as without randomisation. The statistical challenge,
according to Sir David Cox (Cox, 1998), will be to develop methods
of analysis that take account of the complex character of compliance
without analyses becoming too complicated conceptually. He warns
of the dangers of treating compliance as a simple binary, yes, no,
concept and also of ignoring the nature and the essential reason for
noncompliance since this may vary greatly between individuals.
The most extreme form of noncompliance occurs when patients
dropout of a trial prematurely. Dealing with dropouts can present
challenging problems in analysing the results from the trial. The next
section presents a taxonomy of dropouts which will be of importance
in later analysis chapters.

3.3.2. A Taxonomy of Dropouts


The design of most clinical trials specifies that all patients are to have
the measurement of the outcome variable(s) made at a common set
Monitoring Trial Progress 51

baseline
1-

L111

TREATMENT 2
time 1 time 2 time t
nn # I
n

Fig. 3.3. Monotone data pattern caused by patients dropping out of trial.

of time points leading to what might be termed a balanced data


set. But although balanced longitudinal data is generally the aim,
unbalanced data is often what the investigator is faced with be-
cause of the occurrence of missing values in the sense that intended
measurements are not taken, are lost, or are otherwise unavailable.
Staff, for example, may fail to make a scheduled observation, or pa-
tients may miss an appointment. This type of missing value usually
occurs intermittently. In contrast, dropping out of a clinical trial im-
plies that once an observation at a particular time point is missing
so are all subsequent planned observations, giving rise to a monotone
data pattern (see Fig. 3.3).
Assumptions about the probability model for missing data can
influence both the analysis and interpretation of the longitudinal
data collected in clinical trials. In the statistical literature three
types of dropout have been distinguished:
52 Design and Analysis of Clinical Trials

0 Missing completely random (MCAR),


0 Missing at random (MAR),
0 Non-ignorable (sometimes referred to as informative).

To explain the distinction between these three types it is neces-


sary to introduce a little nomenclature:

0 For each patient it is planned to make a sequence of T obser-


vations, Yl,Y2,. . . ,YT. In addition for each patient, there may
be a set of fixed covariates, X , assumed fully observed.
0 Missing values arise from individuals dropping out, so that if
Yk is missing, then so also are Yk+1,.. . ,YT.
0 Define a dropout indicator D for each patient, where D = k if
the patient drops out between the (k-1)th and kth observation
+
time, and D = T 1 if the patient does not drop out.

Completely random dropout (MCAR) occurs when patients


dropout of the study in a process which is independent of both the
observed measurements and those that would have been available
had they not been missing, so that
P ( D = klX,Y,, Yz,.. . ,YT) = P ( D = k)
Here the observed (nonmissing) values effectively constitue a simple
random sample of the values for all study subjects. Examples of
MCAR dropout might be data missing due to accidental death or
because a patient has moved to another district. Intermittent missing
values in a longitudinal data set might also be assumed to be MCAR,
though supporting evidence would usually be required. Completely
random dropout causes least problems for data analyses.
Additionally, data may be missing due to design, but still be in-
dependent of the outcome values. An example would be a study de-
signed to have less frequent assessments in a group having a standard
treatment. Little (1995), distinguishes completely random drop-out
from covariate-dependent dropout, for which
Pr(D = k l X , Yl,. . . ,YT)= Pr(D = k l X )
Monitoring Dial Progress 53

and the probability of dropping out depends on the values of the


fixed covariates X , but given X , it is conditionally independent of
an individual’s outcome values, Y1,.. . ,YT. Such a definition allows
dependence of drop-out on both between-subject and within-subject
covariates that can be treated as fixed in the model. In particular,
if X includes treatment-group indicators, this definition allows the
dropout rates to vary over treatment groups and seasonal dropout
effects could be modelled by including season indicators as within-
subject covariates in the model.
Random dropout (MAR) occurs when the dropout process de-
pends on the outcome measures that have been observed in the past,
but given this information is conditionally independent of all the fu-
ture (unrecorded) values of the outcome variable following dropout,
so that

P ( D = /qX,Yl,. . . ’YT)= P ( D = / q X , Y l , . . . ,Y&1)

Here ‘missingness’ depends only on the observed data with the dis-
tribution of future values for a subject who drops out at time t being
the same as the distribution of the future values of a subject who re-
mains in at time t , if they have the same covariates and the same past
history of outcome up to and including time t. An example of a set of
data in which the dropouts violate the MCAR assumption but may
be MAR is shown in Fig. 3.4, taken from Curran et al. (1998). The
diagram shows mean physical functioning scores by time of dropout;
higher scores represent a higher level of functioning. We can see that
patients with a lower physical functioning tended to dropout of the
study earlier than patients with a higher physical functioning score.
Consequently, the probability of dropout depended on the previous
functioning score and hence the dropout was not MCAR.
Finally, in the case of non-ignorable or informative dropout, the
dropout process, P ( D = k l X , Y1,.. . ,YT)depends on the unobserved
values of the outcome variable. That is, dropout is said to be non-
ignorable when the probability of dropout depends on the unrecorded
values of the outcome variable that would have been observed had
54 Design and Analysis of Clinical Trials

+4 Months
- -
3 Months
B -2 Months
P -%-I Month
c
P

04 4
0 1 2 3 4 Auassmont

Fig. 3.4. Physical functioning score by time of dropout (taken with permis-
sion from Curran et al., 1998).

the patient remained in the study. An example given by Cnaan, Laird


and Slasor (1997) involves trials of patients undergoing chemotherapy
treatment for cancer, in which quality-of-life assessments are required
on a quarterly basis. Most quality-of-life forms are self-report and
may require substantial effort on the part of the patient. Patients
who are experiencing poor quality-of-life are likely to be far less able
to complete the self-report required for response. In this case obtain-
ing valid estimates of population parameters is likely to be far more
complicated since we are in a situation of having to make assump-
tions about the distribution of missing values which cannot be fully
tested by the data.
The full implications of this taxonomy of dropouts for the analysis
of longitudinal data from clinical trials will be made explicit in later
chapters.

3.4. INTERIM ANALYSES IN CLINICAL


TRIALS

The Tuskegee Syphillis Study was initiated in the USA in 1932 and
continued into the early 1970s. The study involved enrollment and
Monitoring Dial Progress 55

follow-up of 400 untreated latent syphilitic black males (and 200 un-
infected controls) in order to trace the course of the disease. In
recent years, the trial has come under severe criticism because of
the fact that the syphilitics remained untreated even when penicillin,
an accepted form of treatment for the disease, became available. Ma-
jor ethical questions arise if investigators elect to continue a medical
experiment beyond the point at which the evidence in favour of an
effective treatment is unequivocal.
Clearly then, it is ethically desirable to terminate a clinical trial
earlier than originally planned if one therapy is clearly shown to be
superior than the alternatives under test. (This may apply even
if it is a different concurrent study which reports such a result.)
But as mentioned in the Introduction to this chapter, in most clin-
ical trials patients are entered one at a time and their responses
to treatment observed sequentially. Assessing these accumulating
data for evidence of a treatment difference large and convincing
enough to terminate the trial is rarely straightforward. Indeed the
decision to stop accrual to a clinical trial early is often difficult and
multifaceted. The procedure most widely adopted is a planned series
of interim analyses to be done at a limited number of pre-specified
points during the course of the trial. Because the data are examined
after groups of observations rather than after each observation, the
name group sequential is often used. A number of such methods have
been proposed, some of which will be discussed in more detail later
in this section. The aim of all the different approaches, however, is to
overcome the potential problems that can arise from repeated tests
of significance or multiple testing.
The problem of taking ‘multiple looks’ at the accumulating data
in a clinical trial has been addressed by many authors including
Anscombe (1954), Armit age et al. (1969), McPherson (1982), Pocock
(1982) and O’Brien and Fleming (1979). The problem is that, if on
each ‘look’ the investigator follows conventional rules for interpreting
the resulting p-value, then inappropriate rejection of the null hypoth-
esis of no treatment difference will occur too often. In other words,
56 Design and Analysis of Clinical Trials

Table 3.2. Repeated Significance Tests on Accu-


mulating Data.

Number of Repeated Tests Overall Significance


at the 5% Level Level
1 0.05
2 0.08
3 0.11
4 0.13
5 0.14
10 0.19
20 0.25
50 0.32
100 0.37
1000 0.53
03 1.o
(Taken with permission from Pocock, 1983.)

repeatedly testing interim data can inflate false positive rates if not
handled appropriately. Armitage et al. (1969) give the actual signif-
icance levels corresponding to various numbers of interim analyses
for a normally distributed test statistic; these values are shown in
Table 3.2. So, for example, if five interim analyses are performed,
the chance of at least one showing a treatment difference at the 5%
level, when the null hypothesis is true, is 0.14. As Cornfield (1976)
comments:

Just as the Sphinx winks if you look a t it too long, so, if you perform
enough significance tests, you are sure to find significance even when none
exists.

An example of this problem described by Freidman, Furberg and


DeMets (1985) involved a trial comparing mortality in clofibrate and
placebo treated patients in the Coronary Drug Project (1981). In the
Monitoring Trial Progress 57

Beta-Blocker Heart Attack Trial


Z-volua 6 -
5 -

4 -

3 -

1 I

1-96

1978
0l Junb 1979
May Ocl.
1979 'Much
1980 % %: Oct.
1991 la2
June i

Data

Fig. 3.5. Results from a trial comparing mortality in clofibrate and placebo
treated patients (taken with permission from Freidman, Furberg and
DeMets, 1985).

early months of the study, clofibrate appeared to be beneficial, with


the significance level exceeding 5% on three occasions (see Fig. 3.5).
However, because of the repeated testing issue, the decision was made
to continue the study and closely monitor the results. The early
difference was not maintained, and at the end of the trial the drug
showed no benefit over placebo. Such a difference might arise for a
variety of reasons, including:

0 early patients in a trial are not always representative of the


later patients,
58 Design and Analysis of Clinical D i a l s

0 number of events are small,


0 randomisation may not yet have achieved balance.
The basic strategy involved in the group sequential approach is to
define a critical value at each interim analysis (Z,(k), k = 1, . . . , K )
such that the overall type I error rate will be maintained at a prespec-
ified level. At each interim analysis, the accumulating standardised
test statistic ( Z ( l c ) , k = 1 , . . . , K ) is compared to the critical value
where K is the maximum number of interim analyses planned for.
The trial is continued if the magnitude of the test statistic is less
than the critical value, for that interim analysis. Various sequences
of critical values have been proposed. Pocock (1977), for example,
suggested that the critical value should be constant for all analyses.
But O’Brien and Fleming (1979) proposed changing critical values
over the K interim analyses. Pet0 et al. (1976) suggested that a large
critical value be used for each interim analysis and then for the last
analysis, the usual critical value should be utilised. Example of the
Pocock, O’Brien and Fleming, and Pet0 et al. boundaries for K = 5
and a = 0.05 (two-sided) are shown in Fig. 3.6. (A brief account of
the type of calculation behind these boundaries is given later.)
For each of the approaches mentioned above the number of in-
terim analyses, K , has to be specified in advance, but Lan and
DeMets (1983) consider a more general method (the alpha spending
procedure) to implement group sequential boundaries that control
the type I error rate while allowing flexibility in how many interim
analyses are to be conducted and at what times.
Pampallona and Tsiatis (1994) introduce a class of boundaries
for group sequential clinical trials that allow for early stopping when
small treatment differences are observed. The boundaries can be
derived exactly for any choice of type I and type I1 error probabilities,
and can be easily applied to both one- and twwsided hypothesis
testing. A brief account of how these boundaries are derived is given
in Table 3.3, while Table 3.4 (adapted from Pampallona and Tsiatis,
1994) gives values of the boundaries for various combinations of the
design parameters. In general, however, the boundaries required for
Monitoring Trial Progress 59

5.8 c

4.0 -
3.0 -

-
L

2.0 -
N-
0
.-
-6
C

$ 1.0 -
-m Accept
E HO
5 0- Continue
z
7J
5 -1.0 -
E
m
z2 -2.0-
v)

-3.0-
-4.0-
/ A Peto

-5.0 /
1 2 3 4 5
Number of Sequential Groups
with Observed Data (i)

Fig. 3.6. Pocock, O'Brien and Fleming and Pet0 et al. boundaries for
interim analyses (taken with permission from Lan and DeMets, 1983).

most proposed group sequential procedures are most easily obtained


using the specialised software described in the Appendix. As an
illustration, we will consider an example from the East manual (see
the Appendix) which involves a trial comparing a new compound
with a standard treatment for the control of systolic blood pressure.
The expectation is that systolic blood pressure should decrease to
about 95 mmHg with the new drug, against the current 105 mmHg
60 Design and Analysis of Clinical Trials

Table 3.3. Obtaining Boundaries for Group Sequential Analysis.

The problem is one of comparing the effectiveness of two treatments, A


and B say.
Let the response be normally distributed with expected values p~ and
p~ respectively, and common variance cr2.
It is required to test the null hypothesis of treatment equivalence, namely
Ho : p~ = p~ = b = 0, against the alternative hypothesis, H1 : S # 0.
Patients will enter the randomised trial in a staggered fashion over time
and the accumulating responses are analysed each time an additional
group of n observations become available on each treatment arm, with
the goal of interrupting accrual into the trial whenever a large treatment
difference is observed.
Let the maximum number of planned analyses be K , so that the maxi-
mum sample size should the trial be required to continue until the last
analysis, is N = 2Kn patients.
The testing strategy consists of stopping the study, and either accepting
Ho or H I , the first time the test statistic takes a value outside a suitably
defined continuation region.
The test statistic at the j t h analysis is defined as:

where X i l A and X i l B denote the responses of the lth patient in the ith
group on treatment A and B, respectively.
The statistic S, is a partial sum of normal random variables, y Z , of the
form:

The group sequential approach entails the specification of a set of critical


values that define appropriate continuation and stopping regions and
that guarantee the desired type I and type I1 error probabilities under
repeated significance testing.
Monitoring Trial Progress 61

Table 3.3 (Continued)


The critical values for early stopping in favour of H I will be of the form
b; = C1 ( a ,p, K , A ) j A , while critical values for early stopping in favour
of Ho will be of the form bj” = jb’ - C2(a,0, K , A)jA, where C1 and C2
are positive constants and by 5 bj’.
C1 and C2 depend on the required significance level a , the size of the
type I1 error p, the maximum number of looks, K , and an additional
parameter A that affects the shape of the continuation region. If A = 0,
the O’Brien-Fleming boundary is obtained. If A = 0.5, the Pocock
boundary results. More generally, Wang and Tsiatis (1987) explore the
family of boundaries to find the value of A that minimises the expected
sample size for various design specifications.
The following stategy can be adopted for a one-sided test

continue the trial if Sj E (by, b:)

At any analysis, values of Sj 2 bj’ will be considered supportive of the


alternative, while values of Sj 5 bj” will be considered as supportive of
the null, and in either case, the trial will terminate.
The value of n can be shown to be

2 ( C 1 + C2)2K2(A-1)
n=2
62

The values of C1 and C2 that satisfy the required operating characteristics


are found using the recursive integration formula described in Armitage
et al. (1969).
(This account summarises that given in Pampallona and Tsiatis, 1993.)

obtained with the standard. From previous experience, the standard


deviation of blood pressure measurements among target patients is
of the order of 15 mmHg. The trial is required to have a power of
90% in order to detect the difference of interest when two-sided sig-
nificance testing is performed at the 5% level. The O’Brien-Fleming
boundary, when the number of looks is set at five, is shown in Fig. 3.7.
This approach requires a maximum of 102 patients; a fixed sample
size design requires 95 patients, but without the possibility of early
stopping.
62 Design and Analysis of Clinical Dials

Table 3.4. Values of C1, C2 and n for various combinations of P,K


and A when S / u = 1 (part of table in Pampallona and Tsiatis,
1994).

1 - ,L3 = 0.80 1- ,L3 =z 0.80


K A c1 c
2 n c1 c
2 n
One sided test, Q = 0.05
4 0.0 3.3118 1.9987 3.52 3.3722 2.7506 4.69
0.1 2.9227 1.7954 3.67 2.9760 2.4440 4.85
0.2 2.6000 1.6225 3.88 2.6454 2.1881 5.08
0.3 2.3395 1.4748 4.18 2.3766 1.9771 5.44
0.4 2.1340 1.3479 4.59 2.1635 1.8043 5.97
0.5 1.9756 1.2383 5.16 1.9981 1.6639 6.71

5 0.0 3.7217 2.2693 2.87 3.7928 3.1052 3.81


0.1 3.2162 1.9973 3.00 3.2775 2.7012 3.95
0.2 2.8023 1.7718 3.19 2.8537 2.3698 4.16
0.3 2.4730 1.5845 3.46 2.5141 2.1021 4.48
0.4 2.2185 1.4279 3.85 2.2502 1.8883 4.97
0.5 2.0268 1.2962 4.42 2.0504 1.7189 5.68

It is, of course, possible that use of some suggested procedures for


interim analyses will lead to inconsistencies. Falissard and Lellouch
(1991), for example, consider a trial planned with four interim analy-
ses and a final one, each analysis occurring after a constant number of
patients in each group. A a test for assessing the difference in treat-
ment means is scheduled for each of the five planned analyses, the
overall type I error required being 5%. Pocock’s method rejects the
null hypothesis if for at least one value of i , i = 1,.. . , 5 , jail 2 2.41.
Now suppose that the results are Izi( < 2.41 for i = 1,... 4 and
a5 = 2.20. An investigator using no interim analyses will reject the
null hypothesis, while one using Pocock’s procedure will accept it.
Thus, the two investigators will reach different conclusions with ex-
actly the same data. Falissard and Lellouch (1991) propose a new
Monitoring Trial Progress 63

approach which eliminates some of these inconsistencies. This re-


quires, for rejecting the null hypothesis, that a succession of T tests
are significant at the current a level. The value of T is chosen so that
the global type I error is also near to a.
If interim analyses are to be part of a clinical trial, the inves-
tigator planning the trial needs to consider both how many such
analyses there should be and how many patients need t o be evalu-
ated between successive analyses. These questions are considered in
detail by F‘ocock (1983) and McPherson (1982). These authors also
provide tables showing power and expected sample sizes for trials
with various numbers of planned interim analyses. For large values
of the treatment difference, the expected sample size is considerably
reduced by many interim analyses. In most trials, however: such
large differences are unlikely, and in these circumstances, Pocock
(1983) suggests that there is little statistical advantage in having a
large number of repeated significance tests. As a general rule, Pocock
recommends a maximum of five interim analyses.
Interim analyses are designed to avoid continuing a trial beyond
the point when the accumulated evidence indicates a clear treatment
difference. As commented above, this is clearly ethically desirable.
But Pocock (1992) suggests that there is a real possibility that in-
terim analyses claiming significant treatment differences will tend to
exaggerate the true magnitude of the treatment effect and that often,
subsequent analyses (where performed) are likely to show a reduc-
tion in both the significance and magnitude of these differences. His
explanation of these phenomena is that interim analyses are often
timed (either deliberately or unwittingly) to reflect a ‘random high’
in the treatment comparison. Simon (1994) also makes the point that
estimates of treatment effects will be biased in clinical trials which
stop early.
Even though group sequential methods can be used to help de-
cide when a trial should be stopped, the subsequent estimation of the
treatment effect and its associated p-value still needs careful corisid-
eration. It is not difficult to find examples of trials in which some
~ ~____ ~~

BWNDIIRl ES
Look Boundary to rdrot
I HO H1
1 4.46 -1.69
2 3.15 -0.06
3 2.57 0.63
4 2.22 1.47
6 1.99 1.99

Fig. 3.7. O'Brien-Fleming boundary obtained from E a s t software for blood pressure example.
Monitoring Dial Progress 65

type of interim analysis was used to stop the trial early, but where the
reported treatment effect estimate and its p-value were not adjusted
for the sequential design but instead calculated as if the trial had been
of fixed size (see, for example, Moertel et al., 1990). Souhami (1994)
suggests that stopping early because an effect is undoubtedly present
may result in a serious loss of precision in estimation, and lead to
imprecise claims of benefit or detriment. Methods that attempt to
overcome such problems are described in Whitehead (1986), Rosner
and Tsiatis (1989), Jennison and Turnbull (1989) and Pinheiro and
DeMets (1997).
It was the Greenburg Report, finalised in 1967 but not published
until 1988, that established the rationale for interim analyses of ac-
cumulating data. In addition, however, it emphasised the need for
independent data monitoring committees to review interim data and
take into consideration the multiple factors that are usually involved
before early termination of a clinical trial can be justified. Such fac-
tors include baseline comparability, treatment compliance, outcome
ascertainment, benefit to risk ratio, and public impact. This type
of committee is now regarded as an almost essential component of
a properly conducted clinical trial and helps to ensure that interim
analyses, by whatever method, do not become overly prescriptive.

3.4.1. Group Sequential Procedures - A


Bayesian Perspective
Any discussion of group sequential procedures would be incomplete
without acknowledging that some statisticians consider the whole
approach almost fatally flawed. Freedman et al. (1994), for example,
find reason for concern in three areas:

0 philosophical awkwardness,
0 how to draw inferences if the stopping rule is not followed,
0 how to estimate treatment effects at the end of a group se-
quential trial.
66 Design and Analysis of Clinical Trials

Freedman et al. (1994) illustrate their philosophical difficulties


with the usual frequentist approaches to interim analyses with the
following (hypothetical) situation:

Suppose that a clinician Dr. C comes to statistician Dr. S for some ad-
vice. Dr. C has conducted a clinical trial of a new treatment for AIDS. He
has treated and assessed 200 patients, and analysis of the results suggests
a benefit from the new treatment that is statistically significant (p = 0.02).
Dr. S, who is also a frequentist, asks how many times Dr. C plans to
analyse the results as the trial progresses. We now consider two possible
responses.
(1) Dr. C may respond that this is the first and only analysis that has
and will be done. He has waited until the data are complete before
analysing them. Dr. S is then prepared to endorse t.he analysis and
p-value.
(2) Dr. C may, instead, respond that he intends to include 1000 patients
in this trial, and that this is the first of five analyses that are planned.
Moreover, the statistician who helped to design the trial had advocated
an O’Brien and Fleming boundary. Dr. S then advises Dr. C that. the
results are not yet statistically significant and that the p-value of 0.02
should not be taken at face value, being one of a series of tests that are
planned.

Freedman et al. (1994) complain that it seems unreasonable that


different inferences should be made by Dr. S depending upon the
plan for further analysis. They then suggest an alternative Bayesian
approach involving the following steps:
0 A prior distribution, representing one’s pre-trial belief about
the treatment difference is specified. (In fact two priors are
recommended, the first t o represent a reasonable sceptic, the
second to represent a reasonable enthusiast .)
0 Data are then gathered during the trial leading to an estimate
of the treatment difference with a confidence interval.
0 Bayes’ theorem is then applied to calculate a posterior distri-
bution that represents one’s current belief about the treatment
difference.
Monitoring Trial Progress 67

0 Recommendations regarding the continuation of the trial are


based upon the posterior distribution. The treatment differ-
ence scale is divided into three ranges:
(1) differences that would lead to a choice of the standard
treatment,
(2) differences that would lead to a choice of t,he new treat-
ment,
(3) an intermediate range in which benefits from the new treat-
ment are balanced by increased toxicity, inconvenience or
cost.

The posterior probabilitites of the treatment effect lying within


each of these three regions may be used to make decisions about the
future of the trial. Freedman et al. (1994) provide an example of such
an approach in a clinical trial investigating the effect of drug combi-
nation of 5-flurouracil and levamisole upon the length of survival of
patients with colorectal cancer.
Although the properties of the Bayesian approach are attractive
in providing an integrated view of all aspects of stopping, it has,
so far, failed to make a major impact on monitoring clinical trials.
Machin (1994) suggests that this may be partly due to clinicians
scepticism over the way that different priors purporting to represent
belief, can influence the interpretation of results. He suggests that if
the Bayesian approach is to evolve into a more than interesting but
unused tool, there is a need for ‘case’ studies to illustrate what it
gives above and beyond current methods. More detailed discussion
of Bayesian methods is taken up in Chapter 9.

3.5. SUMMARY
Once a trial begins, a patient’s progress needs to monitored closely.
Much effort needs to be put into determining whether or not the pa-
tient is complying with the intended treatment and trying, wherever,
possible to ensure that the patient is observed on all the occasions
specified in the trial protocol. The analysis of longitudinal data from
68 Design and Analysis of Clinical Trials

a clinical trial which has a substantial proportion of missing values,


whilst possible with particular statistical techniques, does present
considerably more problems than when the data is complete. The
proportion of missing values in a set of data can often legitimately
be taken as an indicator of the quality of the study.
The standard way of dealing with non-compliance is intention-to-
treat analysis. Only analysis by intention-to-treat can be relied on to
provide an unbiased comparison of the treatment policies as imple-
mented. All other analyses deviate from the principle of randomised
comparison, need to make assumptions that cannot be fully validated
and hence carry a risk of introducing bias. Schwartz and Lellouch
(1967) characterise trials intended for comparing the efficacy of treat-
ment regimens, ‘pragmatic’, and for such trials intention-to-treat is
the only correct analysis.
But in many cases, interest may be more in comparing the drugs
involved in the regimens, what Schwartz and Lellouch label ‘explana-
tory trials’. In this situation, supplementary analyses using com-
pliance information recorded for each patient become acceptable.
Indeed Efron (1998) suggests that taking into account the variable
compliance in a randomised clinical trial may offer advantages such
as the derivation of a dose-response curve for the drugs efficacy, even
though the original experiment was only intended as a single dose.
The basis of the analyses which are eventually applied to the trial
data are the measurements of the outcome variable(s) specified in the
trial protocol. Clearly, the quality of the trial can only be as good
as the quality of the data collected and issues of the reliability of the
chosen outcome measure may need to be addressed.
Continuing a clinical trial beyond the time when there is strong
evidence of a substantial treatment difference is ethically undesirable
and interim analyses which may allow the trial t o be terminated early
are usually written into the protocol of the trial. The statistical-
based boundaries that result are often quite helpful but should not
be viewed as absolute decision rules. In addition, such analyses need
to be used with caution since they can lead to over optimistic claims
about the effectiveness of treatments in some situations.
CHAPTER4

Basic Analyses of Clinical


Trials, the Generalised
Linear Model and the
Economic Evaluation
of Trials

4.1. INTRODUCTION

Senn (1997) makes the point that in its simplest form a clinical trial
consists of a head t o head comparison of a single treatment and a
control in order to answer a single well defined question. The analy-
sis of such a trial might then consist of applying a single significance
test or constructing a single confidence interval for the treatment
difference. In practice, of course, matters tend to be a little more
complex and most clinical trials generate a large amount of data;
for example, apart from the observations of the chosen outcome
variable(s) at different points in time, details of side effects, mea-
surements of laboratory safety variables, demographic and clinical
covariates will often also be collected. As a consequence, the anal-
yses needed may rapidly increase in complexity. In this chapter we
shall restrict attention to methods suitable for trials in which one
or more outcome variables are recorded on a single occasion, usually
the end of the trial. Later analysis chapters will deal with the more

69
70 Design and Analysis of Clinical Trials

involved techniques needed to analyse data from trials in which out-


come measures are observed on several occasions post-randomisation
and possibly also pre-randomisation.

4.2. A BRIEF REVIEW OF BASIC


STATISTICS
The basic statistical principles required in the analysis and interpre-
tation of data from many clinical trials are well described in Pocock
(1983) and elsewhere; consequently, we shall give only a very brief
review of these methods in this section.
Analysis of a single continuous outcome variable observed once at
the end of a trial usually begins with some simple plots and diagrams,
e.g. histograms, and box plots. Figures 4.1, and 4.2 illustrate each
Histog:am for placebo ~ o M t ~ ~ a ~ h

_I
--

Histograms for placebo and active treatment groups in double-


blind trial of an oral mouthwash.
Basic Analyses of Clinical Trials . .. 71

Fig. 4.2. Boxplots for placebo and active treatment groups in double-blind
trial of a n oral mouthwash.

type of plot for the data shown in Table 4.1; these data arise from
a double blind trial in which an oral mouthwash was compared with
a placebo mouthwash. Fifteen subjects were randomly allocated to
each mouthwash and at the end of 15 weeks an average plaque score
was obtained for each subject. This was calculated by allocating a
score of O! 1, 2 or 3 to each tooth and averaging over all teeth present
in the mouth. The scores of 0, 1, 2 and 3 for each tooth corresponded,
respectively to:

0 0 - no plaque,
0 1 - a film of plaque visible only by disclosing,
0 2 - a moderate accumulation of deposit visible to the naked
eye!
0 3 - an abundance of soft matter.

The histogram and boxplot for the placebo group show clear evi-
dence of an outlier, and the corresponding plots for the active group
72 Design and Analysis of Clinical Trials

Table 4.1. Data from Mouthwash Trial.


Subject Plaque Score I Subject Plaque Score

Placebo 1 0.709 Active 16 0.000


2 0.339 17 0.000
3 0.596 18 0.344
4 0.333 19 0.059
5 0.550 20 0.937
6 0.800 21 0.024
7 0.596 22 0.033
8 0.589 23 0.000
9 0.458 24 0.019
10 1.339 25 0.687
11 0.143 26 0.000
12 0.661 27 0.136
13 0.275 28 0.000
14 0.226 29 0.395
15 0.435 30 0.917

indicate the skewness of the distribution. Despite these contraindi-


cations, the data will first be analysed using a two-sample t-test; the
results, shown in Table 4.2, suggest that the active mouthwash has
been successful in lowering the plaque score.
The indication of non-normality for the mouthwash trial data
given by Figs. 4.1 and 4.2, may suggest to some investigators the
need to transform the data in some way before applying the t-test;
alternatively, a nonparametric test of the treatment difference might
be considered appropriate. The question of transformations will be
taken up later, but applying the Wilcoxon rank-sum test gives a
p-value of 0.001, again clearly demonstrating a non-zero treatment
difference (in terms of the two medians). Section 4.5.4 illustrates
another approach for dealing with non-normality that uses bootstrap
estimation.
In many clinical trials, the outcome variable will be binary. An
example is shown in Table 4.3. This arises from a study testing
Basic Analyses of Clinical Trials ... 73

Table 4.2. Results from Applying


Two-sample t-test to Mouthwash
Data in Table 4.1.

Placebo Active
Mean 0.5348 0.2367
SD 0.2907 0.3432
n 15 15
t = 2.567, d.f. = 28, p-value = 0.0159,
95% confidence interval for treatment
difference (0.060, 0.536).

Table 4.3. Aspirin and Cardiovascular Disease.

Myocardial Infarction
Group Yes No Total
Placebo 189 10,845 11,034
Aspirin 104 10,933 11,037

whether regular intake of aspirin reduces mortality from cardiovas-


cular disease. Every other day, physicians participating in the study
took either one aspirin tablet or a placebo. The participants were
blind to which type of pill they were taking.
Such data will often be analysed based on a confidence interval for
the difference in the population proportions of myocardial infarctions
in the aspirin and placebo groups. Details are given in Table 4.4.
The derived confidence interval does not contain the value zero and
suggests that aspirin diminishes the risk of myocardial infarction.
A further measure of the treatment difference often used for bi-
nary outcomes is the odds ratio; its calculation for the aspirin data
is outlined in Table 4.5. The associated confidence interval does not
contain the value one, again indicating the positive effect of aspirin
in preventing myocardial infarction. As we shall see later, the odds
74 Design and Analysis of Clinical Trials

Table 4.4. Constructing a Confidence Interval for the Dif-


ference in Proportions of Myocardial Infarctions amongst
Patients given Aspirin and Placebo.
Of the 11 034 physicians taking placebo, 189 suffered from my-
ocardial infarction over the course of the study, a proportion
pi = 189/11034 = 0.0171.
Of the 11 037 physicians taking aspirin, 104 suffered from my-
ocardial infarction over the course of the study, a proportion
p2 = 104/11037 = 0.00094
The sample difference of proportions is 0.0171 - 0.0094 = 0.0077.
This difference has an estimated standard error of

(0.017)(0.9829) (0.0094)(0.9906)
= 0.0015
11034 +- 11037

A 95% confidence interval for the true difference is 0.0077 &


1.96(0.0015), i.e., (0.005, 0.011).

Table 4.5. Odds Ratio for Aspirin Study.


For the physicians taking placebo, the estimated odds of myocardial
infarction equal 189/10845 = 0.0174.
For the physicians taking aspirin, the estimated odds of myocardial
infarction equal 104/10933 = 0.0095.
The sample odds ratio 8 = 0.0174/0.0095 = 1.832. The estimated
odds of myocardial infarction for physicians taking placebo equal
1.832 times the estimated odds for physicians taking aspirin. The
estimated odds are 83% higher in the placebo group.
The asymptotic standard error of log(8) is calculated as

SE[log(8)] =

=
J' -
1
+-+-
189 10933
0.123
1
10845
1
+-104

Consequently a 95% confidence interval for log8 is log(1.832) *


1.96 (0.123), i.e., (0.365, 0.846). The corresponding confidence
interval for 8 is [exp (0.365), exp (0.846)] = (1.44, 2.33).
Basic Analyses of Clinical Trials . . . 75

Table 4.6. Pain Scores after Wisdom Tooth Extraction.

Method 1 Method 2
Patient ID Age Pain Score Patient ID Age Pain Score
1 27 36 11 44 20
2 32 45 12 26 35
3 23 56 13 20 47
4 28 34 14 27 30
5 30 30 15 48 29
6 35 40 16 21 45
7 22 57 17 32 31
8 27 38 18 22 49
9 25 45 19 22 25
10 24 49 20 20 39

ratio is a fundamental parameter in particular types of models for


binary response variables.
Even in these clinical trials in which comparing two groups with
respect to a single outcome variable is of prime interest, we may still
wish to take one or more other variables into consideration in the
analysis. In particular, if there is a variable known in advance to
be strongly related to the chosen outcome measure, taking it into
account using analysis of covariance can often increase the precision
with which the treatment effect can be estimated. To illustrate a
simple application of analysis of covariance, the data in Table 4.6
will be used. These arise from a randomised trial of two different
methods of wisdom tooth extraction, in which the outcome variable
was a measure of pain on discharge derived from a visual analogue
scale with anchor points, 0 = no pain and 50 = unbearable pain.
The age of each patient was also recorded as it was thought that age
would be related to a patient’s perception of their pain. A plot of
the data is shown in Fig. 4.3.
Details of the analysis of covariance are given in Table 4.7. A
simple regression model is assumed in which, conditional on treat-
ment and age, the pain outcome variable is normally distributed with
76 Design and Analysis of Clinical Trials

Wisdom tooth extraction data

0
L n -

2
2
l 1

2 1

I 1
v
l
.E 52 - 1
a 2 1

2 ' 1
- 7 -
0

2 1 2 2

0
N - , 7
I I I I I I

20 25 30 35 40 45

Fig. 4.3. Scatterplot of age of patient vs. pain score after wisdom tooth
extraction with patients labelled by treatment group.

constant variance. The results of the analysis indicate that pain is


indeed related to age and that the mean pain scores after adjusting
for age differ significantly.
(Particularly commonly used covariates in clinical trials are base-
line, pretreatment values of the main response variable; discussion
of such covariates is taken up in detail in the next chapter.)
Covariates such as age, etc., may also often be of interest when
the outcome variable is binary. But using the linear model given in
Table 4.7 for say, the probability of a zero response ( P ) , immedi-
ately runs into difficulties. The most obvious is that it could lead to
parameter estimates which give fitted probability values outside the
range (0,l). A further problem is that the assumed normal distribu-
tion would now clearly not be realistic. Because of these problems,
Basic Analyses of Clinical T~ials. . . 77

Table 4.7. Analysis of Covariance of Pain Data in Table 4.6.

The analysis of covariance model assumes that pain on discharge and age
are linearly related and that the slope of the regression line is the same for
each extraction method. Specifically the model has the form:

E(pain) = PO + PI group + PZage


where group is a dummy variable taking the value 0 for method 1 and the
value 1 for method 2. The estimated parameter values and their standard
errors are:

variable parameter standard error of estimate/SE


age -0.74 0.23 3.16
group -7.33 3.45 2.13

some suitable transformation of the probability is modelled as a lin-


ear function of the covariates rather than the probability itself. The
most commonly used transformation is the logistic, i.e., X = In -,P
leading to logistic regression. In addition a more suitable distribu-
tional assumption is made.
The linear model used in the analysis of covariance of a normally
distributed response and the logistic regression model for binary
outcomes are unified along with models for other types of response
variables under the generalised linear model to which we now turn
our attention.

4.3. GENERALISED LINEAR MODELS

The binary response variable is an extreme example of departure


from normality for which the linear model described in Table 4.7 is
clearly unsuitable. But in the case of more moderate departures from
this condition, the model has often been applied to some transformed
value assumed to comply with the requirements of approximate con-
ditional normality, linearity and constant variance. Many outcome
78 Design and Analysis of Clinical Trials

measures in common use are, however, resistant to successful trans-


formation. In other cases, the original scale may be a natural one
and effect estimates and inference are more easily understood on this
scale. Measures such as counts of events or symptoms, for example,
typically possess highly skewed distributions where the modal value
is zero or close to it, and do not transform satisfactorily. (Lind-
sey, 1993, draws a distinction between frequency data, obtained by a
classification of sample units typically considered to be independent,
and count data obtained as a response from each sample unit and
typically representing a number of events over a period or features
on a surface.) Such measures can, however, often be readily anal-
ysed by a model in which the log of the expected count is related to
treatment and covariates with observed counts being assumed Pois-
son distributed (or Poisson-like - see later) around this expected
value, a model known as Poisson regression.
Both logistic regression and Poisson regression are special cases
of a wider class of generalised linear models or GLMs. The models
are described in detail in McCullagh and Nelder (1989) and more
concisely in Table 4.8. The essential features are the link and variance
functions; these are detailed for a variety of GLMs in Table 4.9.
Maximum likelihood estimation of the parameters in such models is
fully described in McCullagh and Nelder (1989) but typically consists
of an iterative weighted least squares solution to the score equations,
i.e., the derivatives of the log-likelihood with respect t o the regression
parameters. These equations can be written thus:

U ( P )= c($)ru;l{K
n

i=l
- pz(p)) = 0

where vT1 are weights derived from the variance function vi =


Var(yZ). The large sample covariance matrix of the parameter es-
timates is given by the inverse of the Hessian matrix, H(P), given
by
H(p) =
n

i=l
(%)'ql(%)
Basic Analyses of Clinical Trials . 79

Table 4.8. Components of GLMs.


0 The observed data y1, . . . , y, are assumed t o be a realisation of random
variables Y1, . . . , Y,.
0 The analysis of covariance model described in Table 4.7 is a simple
example of the general linear regression model with p covariates which
can be written as:

where P I , . . . ,PP are parameters that have to be estimated from the


data.
0 For a sample of n observations, the model can be conveniently written
as :
E ( Y ) = xp
where Y’ = [Yl,. . . , Y,], 0’ = [PI,.. . , P P ] and X is the model ma-
trix containing covariate values (usually including a constant t o allow
estimation of an intercept) for the n observations in the sample.
0 Letting p = E ( Y ) , the model and its distributional assumptions can
be written concisely as:

E ( Y ) = p where p = X p

and the response variables are normal with constant variance g2.

0 The model is generalised by first allowing some transformation of the


pis to be a linear function of the covariates, i.e.,

where 77% = g ( p z ) with g(.) being known as the link function. Link
functions are monotone increasing, and hence invertible; the inverse
link f = 9-l is an equivalent and often a more convenient function for
relating p to the covariates.
0 So for a binary variable, g would be the logit function with vi =
log(*) leading to logistic regression. The inverse link is pi = &.
This guarantees that pa is in the interval [0,1], which is appropriate
since pa is an expected proportion or probability in this case.
80 Design and Analysis of Clinical Trials

Table 4.8. (Continued)


0 The second generalisation of the usual Gaussian regression model is an
extended distributional assumption in which the response variables are
assumed to have a distribution in the exponential family, having the
form:
f(?l;0 , 4 ) = e x p w - b ( @ ) / a ( 4 )+ C ( Y , 4 ) )
for some specific functions a ( . ) ,b(.) and c(.).
0 For the normal distribution, for example, 0 = p , q5 = u2 and

0 One aspect of the choice of distribution for a GLM that is fundamental


to estimation and inference is the variance function, V ( p ) ,that captures
how the variance of the response variable Y depends upon the mean,
Var(Y) = q5V(p) where 4 is a constant. For the normal distribution, for
example, the variance does not depend on the mean and V ( p ) = 1,q5 =
u2. For the Poisson distribution the variance is equal to the mean and
VbL) = I-1.

Table 4.9. Parameterisations of Mean and Variance Functions of


Various GLMs.

Normal Poisson Binomial Gamma Inverse Gaussian

Notation N ( p , 0’) P(p) B ( m ,~ ) / m G(p,u ) IG(PL,U Z )


Link

Variance
function 1 P 4 1-P) P2 P3

The validity of the covariance matrix in Eq. (4.2) depends on the


specification of the variance function being correct. An alternative
estimator for the covariance matrix that provides a consistent esti-
mate even when the variance function is specified incorrectly is the
Basic Analyses of Clinical Trials . .. 81

so-called ‘sandwich estimator’, H-’ (p)H-’ ( P ) where

Commonly referred to as a ‘robust’ or ‘heteroscedastic consistent’ pa-


rameter covariance matrix, the use of standard errors derived from
this matrix and the related confidence intervals and p-values can
provide some protection against inadvertent misspecification of the
error component of a model. Such protection is, however, reliant
on asymptotic theory. In small samples and when the model is not
misspecified, this estimator of the parameter covariance matrix can
perform worse than the standard estimator (Breslow, 1990). The r e -
bust estimator also plays an important role in some of the techniques
to be described in the next chapter.
GLMs can be fitted routinely using widely available software -
see the Appendix, and considerably enrich the range of models that
might be applied to the data from clinical trials, in particular mak-
ing it possible to consider appropriate models for outcome measures
with specific properties. Binary variables, for example, are most
often modelled using logistic regression; this estimates effects, and
combines the effects of covariates, on the log-odds scale. Among
other properties, the log-odds scale is symmetric, ensuring equiva-
lent inference regardless of the coding of the binary response. But
when the binary response that is recorded is in fact a truncated count
or an interval censored failure event, with 0 representing absence of
some clinical feature or event, and 1 representing the presence or
occurrence of at least one such feature or event, symmetry may no
be longer appropriate. In such cases, the log-log scale is often more
suitable. This scale is not symmetric but the parameter estimates
relate naturally t o the log-response rate and thus this model has
strong links with methods for survival analysis which are discussed
in Chapter 8.
82 Design and Analysis of Clinical Trials

When binary and count data are presented in grouped form as


y positive responses in a total of m responses, a binomial distribu-
tion might be expected, provided that the m observations are un-
correlated. In some circumstances, however, this assumption may
be suspect. Examples include where the observations are repeated
measures on the same individual (a situation to be considered in
detail in Chapters 5 and 7), or correspond to sets of patients with
each set drawn from, say, one of a number of clinics. In each case
the observations are likely to be correlated. Under such circum-
stances, the variance of the response data is likely to be larger than
that expected from the binomial distribution. There are a number
of ways that such extra-binomial variation or overdispersion can be
accounted for. One is a further generalisation of GLMs in which a
scale parameter 4 is introduced (see Table 4.8), and the resulting
model estimated by a procedure suggested by Wedderburn (1974)
called maximum quasilikelihood. Wedderburn pointed out that the
GLM score equation (4.1) could be solved for any choice of link and
variance f u c t i o n even when their integral (the quasilikelihood) did
not actually correspond to a member of the exponential family nor
even to a known parametric distribution. McCullagh (1983) showed
that the regression estimates obtained from solving the correspond-
ing quasi-score functions were approximately normal, with mean p
and variance still given by Eq. (4.2). McCullagh and Nelder (1989)
propose a simple moment estimator for the scale parameter 4 (see
bottom of Table 4.8) based on the Pearson residuals. (Other possible
approaches to overdispersion are described in Collett, 1991.)
Count data that might usually be modelled by Poisson regression
can also show overdispersion since rates can vary amongst individu-
als. Counts of features, such as metastases or events such as seizures
occurring in individuals, for example, often show extra-Poisson varia-
tion, and consequently are rarely successfully dealt with using simple
Poisson regression. Again, a quasilikelihood approach can be used
to take account of the overdispersion. Alternatively, a more general
parametric model such as the negative binomial can be used.
Basic Analyses of Clinical Trials . . . 83

To clarify the general comments made above and to illustrate


the way in which different choices of model can influence the results,
we now describe a detailed application of GLMs to a set of clinical
trial data.

4.3.1. Models for Counts: the Treatment of


Familial Adenomat ous Polyposis (FAP)
with a Non-Steroidal
Anti-Inflammatory Drug
The example data shown in Table 4.10 come from Giardiello et al.
(1993) and are also reported in Piantadosi (1997). The data relate
to a placebo controlled trial of a non-steroidal anti-inflammatory
drug in the treatment of FAP. The trial was halted after a planned
interim analysis had suggested compelling evidence in favour of the
treatment. Here the longitudinal aspects of the trial will be ignored
(they will be considered in a later chapter) and we shall concentrate
on analysing the count of the number of colonic polyps at 12 months.
A ‘spikeplot’ of the count data shown in Fig. 4.4 shows consider-
able skewness and an extreme value. For this outcome, we will com-
pare the results obtained from an approach using ordinary regression
applied to the log-count, a transformation that substantially removes
the skew, with results from a variety of GLMs, that all use a log-link
but make different assumptions as to how the mean and variance are
related.
A t-test (with group variances assumed equal) for the treatment
difference in log-count gives the value -3.34; the point estimate of
the treatment difference is -1.58 and the associated 95% confidence
interval is (-2.58, -0.58). (This is equivalent to the application of a
Gaussian regression model with treatment group as the only covari-
ate.) Introducing log-baseline count as a covariate again using the
ordinary Gaussian analysis of covariance regression model described
in Table 4.7, gives the estimate and confidence interval shown in the
first row of Table 4.11. These show the expected increase in precision
84 Design and Analysis of Clinical Dials

I
I I I I
1 434
np12
distribution of 12 month polyp count

Fig. 4.4. Spikeplot of count data from a trial of non-steroidal anti-


inflammatory drug in the treatment of FAP.

obtainable from baseline covariate adjustment. The point estimates


in the other three rows of Table 4.11 are roughly comparable, since
they are all estimated on the same log scale. They differ, however, in
terms of the variance functions assumed (see Table 4.9) and this will
be reflected in the standardised residuals, (yi - f i i ) / d m ,ob-
tainable after fitting each model. Probability plots of these residuals
may help in identifying the most appropriate (and least appropriate)
models for these data. Such plots are given in Fig. 4.5. This figure
shows the plots corresponding to the four models fitted in Table 4.11
and, in addition, residuals from a Gaussian regression model for the
raw counts and a standard Poisson regression.
The residuals from the ordinary regression of the raw counts
show the failure of this approach to account for the skewness in
the data, while those from the standard Poisson regression show a
much greater variation than that expected under the assumption that
the polyps represented ‘independent’ events. Uncorrected use of this
Basic Analyses of Clinical Trials . . 85

Table 4.10. Data from Polyposis CTE Clinical Trial (taken with per-
mission from Piantadosi, 1998).

Number of Polyps Size of Polyp


Visit Visit
I D 1 2 3 4 5 1 2 3 4 5 Surg Age Treat

1 0 7 6 3.6 3.4 . . . 1 17 1
2 0 77 67 71 63 3.8 2.8 3.0 2.8 . 1 20 0
3 1 7 4 4 2 4 5.0 2.6 1.2 0.8 1.0 1 16 1
4 0 5 5 16 28 26 3.4 3.6 4.0 2.8 2.1 1 18 0
5 1 23 16 8 17 16 3.0 1.9 1.0 1.0 1.2 1 22 1
6 0 35 31 65 61 40 4.2 3.1 5.6 4.6 4.1 1 13 0
7 0 11 6
1 1 14 2.2 . 0.4 0.2 3.3 1 23 1
8 1 12 7
20 7 16 2.0 2.6 2.2 2.2 3.0 1 34 0
9 1 7 7 11 15 11 4.2 5.0 5.0 3.7 2.5 1 50 0
10 1 318 347 405 448 434 4.8 3.9 5.6 4.4 4.4 1 19 0
11 1 160 142 41 25 26 5.5 4.5 2.0 1.3 3.5 1 17 1
1 2 0 8 1 2 3 7 1.7 0.4 0.6 0.2 0.8 1 23 1
13 1 20 16 37 28 45 2.5 2.3 2.7 3.2 3.0 1 22 0
14 1 11 20 13 10 32 2.3 2.8 3.7 4.3 2.7 1 30 0
15 1 24 26 55 40 80 2.4 2.2 2.5 2.7 2.7 1 27 0
16 1 34 27 29 33 34 3.0 2.3 2.9 2.5 4.2 1 23 1
17 0 54 45 22 46 38 4.0 4.5 4.2 3.6 2.9 1 22 0
18 1 16 10 1.8 1.0 ' . . 1 13 1
21 1 30 30 40 50 57 3.2 2.7 3.6 4.4 3.7 0 34 0
22 0 10 6 3 3 7 3.0 3.0 0.6 1.1 1.1 0 23 1
23 0 20 5 1 1 1 4.0 1.1 0.6 0.4 0.4 0 22 1
24 1 12 8 3 4 8 2.8 1.1 0.1 0.4 1.0 0 42 1

Missing values are indicated by a period.

latter model would have led to grossly exaggerated levels of precision


and significance for the parameter estimates. The residual plot from
the gamma model looks to be the best of the remaining four plots.
More can be learnt about the fit of these four models by con-
sidering a further diagnostic, namely the delta-beta (Ap) measure of
86 Design and Analysis of Clinical Trials

Table 4.11. Results of Fitting GLMs to Polyposis Data.


Model Treatment r-test 95% CI
including baseline Effect
Regression of log-count - 1.42 -4.02 (-2.16, -0.69)
Overdispersed Poisson -1.43 -4.69 (-2.03, -0.83)
Gamma -1.25 -5.11 (-1.74, -0.77)
Inverse Gaussian -1.26 -4.54 (-1.81, -0.72)

influence proposed by Pregibon (1981) and Williams (1987). This is


an example of a single case deletion diagnostic that measures the im-
pact of a particular observation on a specific parameter in the model.
In general, it is the standardised difference between the parameter
estimates with and without a particular observation included. Here
our interest will centre on the Aps associated with the estimated
effect of treatment on the log-count of polyps. (The diagnostic is
calculated simply by repeated model estimation with each case in
turn excluded.)
Index plots of the Aps for the four models fitted to the polyposis
data are shown in Fig. 4.6. The first row of numbers at the top of this
plot gives the baseline polyp count used as one of the covariates in the
analysis and the second row gives the observed count at the end of 12
months. It is clearly the observation with very low outcome count,
namely observation 23 with an outcome of 1, that is most influential
in the approach using ordinary regression on the log-counts. Such
a finding is not uncommon and cautions against the unquestioning
use of the log( l+count) transformation frequently used when zero
counts are encountered. For the log-link models, however, this data-
point is relatively less influential. Among the log-link models, which
points are influential can be understood by a consideration of the
mean-variance relationships implied by the choice of error distribu-
tion. As can be seen from Table 4.9, the Poisson model assumes the
variance to increase in proportion to the mean; for the gamma it
increases with the mean squared, while for the inverse-Gaussian it
1w-l 0 ) 1 00 * t

O M O M
E:
0251 / 025- 0 25

I o w 0 I o w 000000
o w 025 o m 075 < w OD0 025 O M 075 100 O W 025 O m 075 100
Emplrlul PIIJ - IIIN-1) Emplrlul PII) - UINI1) Emplrlul PI11 - II(N11)
Ordinary regression of count Ordinary regression of log-count Poisson GLM with log-link

c
~~~~~
O M 0 1
505

1 025
z
O W O W
000 025 O M 075 1w 000 015 O M 075 100
Emplrlul PI11 - IIIN+1) I
I n w l r l ~PI11 - u(N11)
Overdispersed Poisson GLM with log-link Inv Gausian GLM with log-link

Residual plots for Polyposis Data


Fig. 4.5. Probability plots of residuals from fitting four models to the polyposis data.
88 Design and Analysis of Clinical Trials

log.nomal regression 0,inv. Gaussian x


Overdispersed Poisson +, Gamma 0
7 5 23 35 11 12 7 318180 8 20 11 24 34 54 30 10 20 12

4 26 16 40 14 16 11 434 26 7 45 32 80 34 38 577 1 8
.25
t

X 8
e x X t

'0 i
84
N
CI
O
i t
pl X
nI
x o t

.m
0 0
CI

0
0
X
-.25
t

-.5
I I I I I I
4 8 12 16 20 24
id
Treatment delta-beta's for polyposis
Basic Analyses of Clinical Trials . . . 89

increases with the mean cubed. For these data the gamma model,
with variance increasing with the mean squared, would seem most
appropriate with all the observations having similar and low influ-
ence. For the Poisson model, the expected variance of the points
with the highest mean (those with the highest baselines) is too small,
with these points being attributed greater influence. For the inverse-
Gaussian model, the expected variance of those with the lowest
means (the lowest baselines) is too small, and correspondingly these
points are found to have greater influence.
An alternative way of tackling the problem of non-normal condi-
tional errors using bootstrap is considered in Section 4.5.

4.3.2. Ordinal Response Models: A Clinical


Trial in Lymphoma
The data shown in Table 4.12 arise from a clinical trial of cytoxan+
prednisone (CP) and BCNU+prednisone (BP) in lymphocytic lym-
phoma. The outcome variable is the response of the tumour in each
patient, measured on a qualitative scale from ‘complete response’
(best) to ‘progression’ (worst). It is plausible that this scale repre-
sents a unidimensional ordinal scale. A variety of models will now
be considered for these data.

Table 4.12. A Clinical Trial in Lymphoma.


BP CP Total
Complete response 26 31 57
Partial response 51 59 110
No change 21 11 32
Progression 40 34 74
Total 138 135 273

4.3.2.1. Models with parallel linear predictors


Two natural extensions of the binary logistic regression model that
might be considered for these data are the proportional-odds model
and the continuation-ratio model.
90 Design and Analysis of Clinical Trials

The proportional-odds model can be considered as an extension


of the binary logistic model in which, instead of a single threshold
partitioning scores on a continuous dimension into two categories,
multiple thresholds result in multiple partitions. The log-odds of the
probability of falling in any category to the right (1- & < k p i j ) and
left ( c j < k p i j ) of the binary partition, formed by considering each
threshold k in turn, is assumed to conform to a model with parallel
linear predictors, such that for individual i:

This construction of model fits naturally in circumstances in which


pooling of adjacent categories could be plausibly considered. This
model is one of many that are based on categories defined by the po-
tentially arbitrary partitioning of a cumulative distribution function.
Conceptually, very similar models using the complementary-log-log
and probit links are possible.
The continuation ratio model is an alternative model that ex-
tends the ordinary logistic model by considering the categories its
occurring sequentially, with each advancement to the next category
conforming to binary logistic models with parallel linear predictors,
but where advancement to the next category is conditional upon
having advanced through all previous categories. For this model:

Again, complementary log-log or probit link functions could be used


instead of logistic.

4.3.2.2. Non-parallel models


In the preceeding models, only the intercept a varied with the cat-
egory; the regression coefficients were assumed common. Allowing
Basic Analyses of Clinical Dials .. . 91

the regression coefficient to vary with k allows for non-parallel regres-


sions, with effects that apply at some levels but not others. In many
applications, the question as to whether parallel linear predictors can
be assumed may be an open one. For example, the biochemical action
of some drug treatment might be thought likely to interrupt disease
progression at one stage transition (threshold) rather than another.
The ability to test for heterogeneity in regression coefficients across
thresholds may therefore be important. Models with such flexibility
are sometimes referred to as partial-proportional models (Peterson
and Harrell, 1990).

4.3.2.3. Model estimation


The models described above can be estimated by direct maximum-
likelihood using purpose written programs. However, both parallel
and non-parallel models can be estimated by forming multiple records
for each subject, each record representing the response on each of the
notional component binary response models described above. In the
case of the proportional odds models and similar models based on
the other possible link/cumulative distribution functions, the records
from the same individual cannot be considered as independent (Snell,
1964; Clayton, 1974)) but instead form a single multivariate set.
There are a number of more and less efficient ways in which this
correlation can be accounted for, some of which are considered later.
The advantage of this overall approach is that tests of the signifi-
cance of regression terms involving the interaction between dummy
variables for record and predictor variables can be used to test for
non-parallel linear predictors.
Figure 4.7 shows the cumulative distribution functions over the
ordered response categories for the two treatments for lymphocytic
lymphoma. This simple plot suggests there to be rather little dif-
ference in the two treatment groups. Fitting an ordinal logistic re-
gression model to these data using maximum likelihood, gave an
estimated log-odds ratio of 0.322 in favour of the CP treatment but
92 Design and Analysis of Clinical Trials

0
I I I I
0 1 2 3
best-worst
CDF of qualitative response from lymphoma trial

Fig. 4.7. Cumulative distribution functions over the ordered response cat-
egories for the two treatments of lymphotic lymphoma.

with a 95% confidence interval from -0.112 to 0.756. The estimate


obtained from fitting the same model but specified as a set of three
binary logistic regressions (with responses complete versus the rest,
complete or partial versus the rest, and the rest versus progression)
with different intercepts but common slope, and under the assump-
tion of independence, gave an estimate of 0.314. Of course, the
standard errors from such a model will be incorrect, since the three
responses will be correlated.
One approach for dealing with this correlation turns out to be
remarkably straightforward and will be one that we turn to repeat-
edly throughout this book. We have already introduced the het-
eroscedastic consistent/sandwich/robust variance estimator earlier
in this chapter. This gave consistent estimates of standard errors
even when the chosen model error was inappropriate. The central
part of the sandwich defined in Eq. (4.3) assumes independent ob-
servations. In our correlated multivariate response context, this can
Basic Analyses of Clinical Trials . . . 93

be achieved simply by summing the score contributions within each


cluster to form ‘super-observations’ that are independent. The stan-
dard formula can then be applied (Binder, 1983; Rogers, 1993).
Using this robust covariance matrix approach for testing corre-
lated multivariate responses allowed a test for non-parallel linear pre-
dictors (non-proportionality in this case), corresponding to whether
treatment effects were uniform across all levels of disease severity.
The Wald chi-square of 0.10 (p = 0.9) obtained for this 2-degree-of-
freedom interaction between treatment and the three severity thresh-
olds, provided no evidence for non-parallel effects. There was little
evidence for the treatment effect varying with severity.
A further example of the application of a generalized linear model
will be given in Chapter 5.

4.4. MULTIPLE ENDPOINTS

While a small minority of large simple trials may focus on a single


outcome measure, for example, mortality, in most circumstances this
would be an oversimplification of the diversity of patient response.
In many disease conditions, response to treatment can have many
different aspects; consequently, clinical trials frequently lack a sin-
gle definitive outcome measure that completely describes treatment
efficacy. When a treatment is thought to affect a disease in a mul-
titude of ways, several outcome variables may be necessary to fully
describe its effect on patients. Berkey et al. (1996) give an exam-
ple involving the efficacy of second-line drugs in rheumatoid arthritis
(i.e., drugs used after the initial standard therapies are unsuccess-
ful). Efficacy is evaluated on a variety of measures, often including
tender joint count, erythrocyte sedimentation rate and grip strength.
In general, the multiple outcomes might include clinical events, symp-
toms, physiological measurements, blood tests, side effects and qual-
ity of life.
Pocock (1996) suggests that one possible approach would be the
pre-specification of priorities amongst the outcome measures. This
94 Design and Analysis of Clinical Trials

would have the result of providing a clear framework for emphasis


(and de-emphasis) of results in eventual publication. Such prioriti-
sation is at its simplest when a single outcome measure is selected
in advance. As Pocock points out, however, when results actually
arrive, it can be difficult to adhere t o such principles.
Comparing treatment groups on each of the outcome measures
at some chosen significance level a will, as we have seen earlier in
this chapter, inflate the type I error. In the unlikely event that the
measures are independent, the probability of declaring that at least
one is significant when there are in fact no treatment differences is:

P = 1 - (1 -

where m is the number of outcome measures. For cy = 0.05 this leads


to:

m P
1 0.05
2 0.0975
3 0.14625
4 0.18549
5 0.22622
10 0.40126
20 0.64151
50 0.92306

Various procedures have been proposed for keeping the proba-


bility that we reject one or more of the true null hypotheses in a
set of comparisons (the fumilywise error, FWE) below or equal to a
specified level a. The most familar of these is the Bonferroni pro-
cedure which controls the FWE rate by conducting each test on an
outcome measure at level a/m. This is simple to apply but suf-
fers from considerable drawbacks. The first is that it is excessively
Basic Analyses of Clinical Trials . . . 95

conservative, particularly if m is large. Consequently, the rather un-


satisfactory situation can arise where many tests are significant at
level a but none at level a/m. In addition, the Bonferroni correc-
tion ignores the degree to which the chosen outcome measures may
be correlated, which again leads to conservatism when such correla-
tions are substantial (see Blair et al., 1996). There are alternative
approaches which, partially at least, overcome the first problem -
see, for example, Holm (1979) and Hochberg (1988). Unfortunately,
these alternatives do nothing to address the second problem of cor-
relations between measures. More recently, Blair e t al. (1996) have
described permutational methods which appear to overcome the con-
servatism problem even in situations with many outcomes which have
large correlations.
Even if some appropriate adjustment of p-values could be found
there remains the risk of data dredging and distortive reporting by
a post hoc emphasis on the most statistically impressive findings.
An alternative to testing each outcome measure separately at a
significance level calculated to control the FWE rate, is to use a
procedure which simultaneously tests for treatment differences on all
variables. Such tests take into consideration the empirical correla-
tion structure of the outcome measures, thus overcoming one of the
criticisms of methods such as the Bonferroni correction. The most
commonly used of these global tests is Hotelling’s T2,which is the
multivariate analogue of the two-sample t-test. To test the equality
of two mean vectors against the alternative that they differ in some
respect, the T2 statistic is:

where n1 and n2 are the number of observations in the two groups and
n = n1 +n2; 21 and 2 2 are the sample mean vectors in each group and
S is the estimate of the assumed common covariance matrix given
bv:

where S1 and S2 are the sample covariance matrices in each group.


96 Design and Analysis of Clinical Trials

Under the equality hypothesis and assuming that the outcome


measures have a multivariate normal distribution with the same co-
variance matrix in each of the two treatment groups, then:
F = ( n - m - 1)T2/(n- 2)m (4.8)
has an F distribution with m and n - m - 1 degrees of freedom.
As an example of the use of Hotelling’s test, it will be applied to
the size and log-count outcomes at 12 months in the polyposis data
given in Table 4.10. The results are detailed in Table 4.13. The test
indicates a significant treatment difference on the bivariate mean of
the two variables.

Table 4.13. Hotelling’s Test for Log-count and Size Variables


at 12 Months in Polyposis Data.
0 The variance-covariance matrices of the two treatment groups are;

s1 = ( 1.001
0.4903 0.5321
0.4903

sz = ( 1.1298 1.1956
1.1956 1.9975

0 The combined estimate of the assumed common covariance matrix


is

s= ( 1.0162 0.8222
0.8222 1.2217

0 The sample mean vectors are

Xi = [3.7526,3.1100]

XL = [2.1726,1.1833]

0 These lead to T 2 = 11.1612 and F = 5.2523. The associated p-value


is 0.018.
Basic Analyses of Clinical Dials . . . 97

Although Hotelling’s T 2 test accounts for the correlations be-


tween the outcome measures, it is not without its own problems.
First simultaneous testing of all the outcome variables may not an-
swer a practically relevant problem; it is only really useful when the
different variables measure different aspects of the same underlying
concept. Second the ‘directionless’ alternative hypothesis associated
with the test may mean that it lacks power. Some alternatives which
are more powerful if treatment improves all outcomes are described
in Follman (1995).

4.5. ECONOMIC EVALUATION OF


TRIALS

Of increasing importance to health service providers are questions


that go beyond determining whether new treatments are more effec-
tive than the current standard treatment. Among these are questions
relating to treatment costs. For example, at its simplest, where a new
treatment is quite obviously cheaper than the current standard, then
a trial that shows that a new treatment is equivalent in effectiveness
to the standard may be sufficient reason to argue for the adoption of
the new treatment over the old.
The appeal of economic evaluations of trials is at least two-fold.
First, policy makers and accountants immediately find the results of
research more interesting and relevant, and thus may be persuaded
by them. Secondly, the approach seems to take a disparate set of
input and outcome measures and sum them up on a single ‘objec-
tive’ scale, one measured in monetary units. Considerable caution
is required on both counts to ensure that real clinical and patient
benefits are not overlooked and that the unit costs used a s weights
in this economic weighted sum are indeed appropriate.

4.5.1. Measuring Costs


In practice routine audit systems rarely give adequate data for a
proper evaluation of costs. Thus in practice trials may need to
98 Design and Analysis of Clinical Trials

include extended measurement protocols that can provide full treat-


ment costs at the individual patient level. A variety of issues need
to be borne in mind when considering these measures:
(1) Defining a cost is surprisingly complex. Are costs borne by pa-
tients rather than treatment providers to be included? Time off
work might not be, but what about the costs of travel to receive
treatment, or costs that are passed on directly from treatment
provider to patient?
(2) Some costs are transfer costs and should not be included. Differ-
ent treatments may involve a transfer in the billing of the same
cost from one department to another, with no net change in cost.
(3) Costs should not be included for services that would not have
been used elsewhere. For example, consider a new treatment
that makes use of some currently rarely used but nonetheless
necessary piece of equipment. Provided the use made of this
equipment by the new treatment does not conflict with its cur-
rent use, then much less than the full cost of this equipment
should be attributed to the new treatment. The calculation of
the appropriate amount, the sc-called marginal opportunity cost,
is often far from straightforward.
(4) Costs should not be counted twice. Thus drug costs charged to
patients should not be included in both hospital costs and patient
costs.
(5) In the same way that clinical outcomes are monitored and com-
pared for a specified period of time, so too it is the case for costs.
Longer term treatment benefits might include lower use, or some-
times greater use, of quite a range of health service facilities for
complaints not obviously directly related to that treated. For
example, patients with successfully treated heart conditions may
experience longer term psychiatric problems that are costly to
treat. Are these costs to be included? A further complication
is that in most branches of economics it is usual to apply a dis-
count rate to future costs, a reflection of the fact that where costs
are deferred interest can be earned on the corresponding funds.
What, if any, discount rate should be applied?
Basic Analyses of Clinical Trials . . . 99

As a consequence of issues such as these, costs are easier to define


within a small closed economic unit than within a community as a
whole and can be substantially different. The cost of a treatment as
viewed from the perspective of a single private health care provider
can be very different from that viewed from the perspective of a
national health service.
The uncertainties as to the inclusion criteria and amounts to as-
sign makes this area of measurement one that should be subject
to the same rigours as the rest of the trial protocol. This should
include the need to define the range of eligible costs prior to randomi-
sation; the need for blindness in the collection of the economic data;
and, particularly where the determination of unit costs are part of
the analysis stage (i.e., are not known and agreed prior to the study),
the importance of blindness and probably also ‘independence’ at the
dat a-analysis st age.

4.5.2. Analysis of Cost Data


When it comes to analysis, further considerations need to be borne
in mind that relate to how the costs come about.

(1) The first relates to the characteristics of the treatment. In many


cases, although treatments are routine, and thus might be con-
sidered to have a roughly standard fixed cost, they are routine
only for a proportion of patients. For the remainder, a variety of
complications and side effects result in the need for repeated, dif-
ferent or additional treatment. As a consequence, the distribution
of costs by patient is typically highly skewed, often with a small
number of patients accounting for a disproportionate amount of
the costs.
(2) Economic costs are a measure where the scale of measurement is
fixed and known. As a consequence, analyses that attempt to deal
with the non-normality of the cost outcome by transformation are
not appropriate. This is because the estimated difference in log-
costs (or any other nonlinear transformation) of two treatments is
100 Design and Analysis of Clinical Trials

not the same as the log of the estimated cost difference. The anal-
ysis of transformed costs simply does not answer the question of
interest. Instead, an approach that analyses the untransformed
cost using some appropriate GLM should be used, for example,
using a model with gamma or inverse-gaussian distributed errors
and log-link function. In such a model, the use of the log-link
function means that treatment and other effects are estimated in
terms of a multiplicative effect, implying an ability to report find-
ings in terms of a percentage reduction or increase in mean costs.
Although slightly less familiar to accountants than reporting an
absolute cost difference (which can anyway still b e derived), the
structure of the model recognizes that costs cannot fall below
zero. An alternative approach is to persist in the use of OLS
linear model methods that give consistent estimates regardless
of the distribution of costs, but to recognize their non-normality
by the use of a method such as bootstrap resampling to estimate
the standard errors. A third approach that may be more helpful
when the cost distribution shows signs of bimodality, is to fo-
cus attention more directly on the factors, including treatment,
that distinguish the high cost from the standard cost individuals,
using logistic regression on a binary split of the cost data.
(3) Almost always patient costs are derived by the application of unit
costs to data on the number of units- used by each patient. The
units might include days and nights on an in-patient ward, num-
ber of outpatient assessments, units of blood products infused
and so on. At least two potentially important consequences fol-
low. Firstly, if there is variation in the costs of the same units
between patients it is unlikely that. this variation occurs at the
patient level, but more often at the level of the medical centre,
supplier or some such. Thus, from the point of view of costs,
patients may fall within a much more complex sampling design,
perhaps with nesting within centre. or within a crossed design
of suppliers of different products or services. A failure to recog-
nize such clustering may give a misleading impression as to the
Basic Analyses of Clinical Trials . . . 101

precision in the estimates of costs and cost differences. Secondly,


and perhaps more importantly, there is typically considerable
uncertainty in the costs of many of the units measured, and since
the same unit costs are commonly applied across many or even
all of the patients, a different choice of unit cost can substantially
alter the overall results of a study. A common response to this
problem is to narrow the focus of the analysis merely to those
costs that can be well measured. Sometimes such costs represent
a trivial proportion of the total costs and to narrow the focus in
this way then makes very little sense. An alternative response
is to consider a range of values for each unit cost, presenting
the results in the form of a sensitivity analysis. This typically
provides results of little value, since it is frequently the case that
few differences prove to be robust under the whole of the plausible
unit cost space. One sensible way to approach this issue would
be to formulate sensible distributions for unit costs and then
to use these within a simulation-based estimation method (see
Chapter 9).

4.5.3. Cost-Benefit Analysis


The previous section makes clear that in the presence of uncertainty,
the analysis of cost data from a trial may be far from straightfor-
ward. In practice, the economic evaluation of benefits are still more
complicated and uncertain than those of costs. Benefits may accrue
not only to patients but to patient carers, their partners and families,
and to the wider society. Deciding where to draw the line, the placing
of monetary values on benefits that may be of such very disparate
kinds, and the social equity issues in deciding whose valuations they
should be, has persuaded most not to attempt such a project.

4.5.4. An Example: Cognitive Behavioral


Therapy
The RCT of cognitive-behavioral therapy for psychosis by Kuipers
et al., (1998) is a good example of recent attempts to incorporate
Design and Analysis of Clinical Trials

an economic perspective into trial evaluation. A subset of the raw


data is presented in Table 4.14. In view of the likely novelty of such
measures to statisticians, we describe this study in some detail. As
mentioned above as a currently common occurrence, the economic
evaluation of this trial was not part of the original protocol. Costs
during the primary nine-month treatment phase of the study were ob-
tained by applying estimated unit costs to in-patient hospital service
utilisation records. This necessarily represented a restricted range
of items within the wider cost space that might have been exam-
ined, essentially focussing on the direct service costs of the CBT
treatment itself over standard treatment. The 9-18 month follow-up
phase was more systematically evaluated using a Client Service Re-
ceipt Inventory (CSRI; Beecham and Knapp, 1992). This considered
a wider range of costs, including some accomodation costs, but did
not include informal care-giver support nor potentially illness-related
unemployment. Again, cost figures were commonly derived by the
application of unit costs, often based on adjustment of national esti-
mates of long-run marginal costs (Netten and Dennett, 1996) rather
than actual costs (the latter may have been neither known nor nec-
essarily appropriate).
In common with most community-based trials of severe psychi-
atric disorder, loss to follow-up represented a considerable problem,
made worse by additional non-response to the CSRI, the economic
evaluation instrument. The figures of Table 4.15 relate to the 31 out
of 60 subjects for whom full cost data could be calculated. Figure 4.8
presents box-plots of the total cost for each treatment group, showing
the distributions to have the expected heavier upper tails. Various
methods for calculating the cost differential and its confidence inter-
val are shown below. Among these is bootstrap estimation (Efron,
1979) in which the precision of estimators is examined by using their
empirical distribution when repeated on many resamplings from the
sample actually observed. The new samples are of the same size, but
the sampling is done with replacement, and consequently the sample
members change from sample to sample. Confidence intervals for
Basic Analyses of Clinical Trials . . . 103

Table 4.14. Estimated Patient Costs from the CBT Trial.


(Kuipers et al. 1998 and pers. comm.)

Estimated Costs
Group Accomod’n Followup Treatment Total
1. cbt 4953 630 5583 11166
2. cbt
3. cbt 5589 12372 18261 36522
4. cbt 6084 2177 8261 16522
5. cbt 3120 2387 5507 11014
6. cbt
7. cbt
8. cbt 4680 1044 5724 11448
9. cbt
10. cbt
11. cbt 5499 21599 27098 54196
12. cbt 6084 1517 7601 15202
13. cbt 6162 11460 17622 35244
14. cbt 7917 352 8269 16538
15. cbt
16. cbt 8814 270 9084 18168
17. cbt
18. cbt 6162 438 6600 13200
19. cbt 7917 1067 8984 17968
20. cbt 7137 15329 22466 44932
21. cbt
22. cbt 3120 3811 6931 13862
23. cbt 4953 109 5062 10124
24. cbt 7683 1670 9353 18706
25. cbt
26. cbt 3212
27. cbt
28. cbt 8814 208 9022 18044
29. cbt
30. control
104 Design and Analysis of Clinical Trials

Table 4.14 (Continued)

Estimated Costs
Group Accomod’n Followup Treatment Total
31. control
32. control 6162 6645 12807 25614
33. control 6084 852 6936 13872
34. control
35. control
36. control 6786 6015 12801 25602
37. control 11349 4047 15396 30792
38. control
39. control 5889 2016 7905 15810
40. control 3120 1742 4862 9724
41. control 6786 4734 11520 23040
42. control 4953 1026 5979 11958
43. control
44. control 5889 12323 18212 36424
45. control
46. control
47. control 6786 1600 8386 16772
48. control
49. control
50. control 6162 738 6900 13800
51. control 7917 2334 10251 20502
52. control 7059 25383 32442 64884
53. control
54. control
55. control 6084 21364 27448 54896
56. control
57. control
58. control
59. control 7137 440 7577 15154
60. control
Basic Analyses of Clinical Dials . . . 105

Table 4.15 Economic Evaluation Of Cognitive Therapy Trial.


Cost Difference
UK-pounds 95% CI
(CBT-control) CI
Simple t-test (equal variances) -3912 (-14407,6583)
Simple t-test (unequal variances) -3912 (-14588, 6765)
Robust GLM (overdispersed gamma) -3912 (-13782, 5959)
Bootstrap linear regression -3912
(1000 samples)
normal approximation method (-13917, 6094)
empirical percentile method (-14002, 5307)
bias-corrected percentile (-14710, 4962)

8 Total cost - CBT 8 Total cost - control


64884 0

1 0
-

0
0

-
1
9724 -
106 Design and Analysis of Clinical B i a l s

a parameter estimate can be based on a normal approximation using


an estimate of the standard deviation of the empirical distribution,
or can be based on the quantiles of the empirical distribution (re-
quiring more samples but not assuming normality). These can also
be adjusted to yield so-called bias corrected intervals (Efron and
Tibshirani, 1986). Both the GLM and bootstrap percentile meth-
ods generate asymmetric intervals reflecting the skew in the cost
distributions.
As in many studies some unit costs were easier to measure than
others and some represented a substantial proportion of the total
cost. In this study, the number of days of inpatient care was relatively
easily obtained, sometimes even when patients were non-responders
at follow-up. The simple correlation between this variable and total
cost was 0.93. Thus this variable by itself might be considered as
a useful surrogate for overall cost, with the advantage that it was
available for 53 patients rather than just the 32 with total cost data.
Analysis of this variable on the larger sample supported the lack
of treatment difference in costs, the bootstrap percentile interval for
the difference in in-patients days being -32.0 to 7.10. An alternative
use for this variable would be to use it to reweight the sample with
complete cost data in an attempt to correct for selective loss. This
issue is taken up in later chapters.

4.6. SUMMARY

The primary data for analysis in many clinical trials will consist of
the observations of one or more outcome variables taken at the end of
the trial. In many cases the analysis might consist of the construction
of a simple confidence interval appropriate for the type of variable at
hand. When covariates other than treatment group are of interest,
then it will usually be necessary to consider some form of model for
the data. Many investigators often use Gaussian-based regression for
all but binary variables, where simple logistic regression is usually
applied. But it will frequently be advantageous to consider more
Basic Analyses of Clinical Trials . . . 107

carefully the nature of the outcome variable and try to find a more
suitable GLM.
One important element in this choice of model is the scale upon
which it is desired to estimate effects. For economic data, the scale
is clear and one for which a transformation of the response variable
would not be helpful. Appropriate choice of GLM or the use of meth-
ods such as bootstrap that are based on the empirical distribution
were shown to be more suitable. Unlike most clinical measures, the
construction of economic measures draws on a considerable amount
of information external to the trial. Trial protocols should where
possible bring the construction of these measures within their scope.
The common practice of using unit costs may also generate complex
patterns of correlated measurement error not easily dealt with at the
analysis stage.
When multiple endpoints are available a variety of methods might
be applied. If the number of outcomes is small, say less than five, then
adjusting significance levels using the Bonferroni correction or some
less conservative adjustment may suffice. Alternatively, simultane-
ous inference using Hotelling’s test may be required. With a large
number of outcome measures the problems become more difficult.
In many cases having many outcome measures may be indicative of
a poorly thought out study. Pocock (1996) points out the merit in
drawing up pre-defined strategies for statistical analysis and report-
ing of trials with appropriate predeclaration of priorities, since there
is a clear need to safeguard against manipulative post hoc emphases
and distortions in conclusions. On the other hand, a protocol that
is too prescriptive may lead to inflexibility and a supression of both
clinical and statistical ‘insight’.
CHAPTER
5

Simple Approaches
to the Analysis of
Longitudinal Data
from Clinical Trials

5.1. INTRODUCTION

Medical treatments rarely result in a one time final result for a pa-
tient; generally, they require clinicians to follow the evolution of a
patient’s health over a period of time. Consequently, in the ma-
jority of clinical trials the primary outcome variables are observed
on several occasions post randomisation and often also prior to ran-
domisation. Such longitudinal data can be analysed in a variety of
ways. In the last decade, many powerful new methods have been
developed which allow a variety of potentially useful models to be
applied, many of which can be employed when the data are unbal-
anced for whatever reasons. (As we shall see in the next two chapters,
however, fitting such models to unbalanced data requires consider-
able care if misleading inferences are to be avoided.) Some of the
newly developed techniques are also able to deal with non-normal
outcome variables, e.g., binary variables indicating the presence or
absence of some characteristic. This methodology will be considered
in detail in Chapters 6 and 7.

108
Simple Approaches to the Analysis of.. . 109

In this chapter, however, we concentrate on an account of a num-


ber of relatively simple approaches to the analysis of longitudinal
data. For many studies these methods may provide a perfectly ade-
quate analysis, for whilst simple, they are not necessarily simplistic
and are often extremely useful when applied sensibly. Even when
they fail to provide the complete answer to the analysis of a set of
longitudinal data, they may frequently prove to be useful adjuncts
to the more complex modelling procedures to be described later. As
with most data analysis situations, graphical displays of the data are
an essential preliminary step and this is where we begin.

5.2. GRAPHICAL METHODS FOR


DISPLAYING LONGITUDINAL DATA

Graphical displays of data are often useful for exposing patterns


in the data, particularly when these are unexpected; this might be
of great help in suggesting which class of models might be most
sensibly applied in the later more formal analysis. According to
Diggle et al. (1994), there is no single prescription for making effec-
tive graphical displays of longitudinal data, although they do offer
the following simple guidelines:

0 show as much of the relevant raw data as possible rather than


only data summaries;
0 highlight aggregate patterns of potential scientific interest;
0 identify both cross-sectional and longitudinal patterns;
0 make easy the identification of unusual individuals or unusual
observations.

A number of graphical displays which can be useful in the pre-


liminary assessment of longitudinal data from clinical trials will now
be illustrated using the data shown in Table 5.1. These data arise
from a double-blind, placebo controlled trial involving 61 women with
major depression, that had begun within 3 months of childbirth and
110 Design and Analysis of Clinical Trials

Table 5.1. Data from Trial of Oestrogen Patch for Treating


Post-natal Depression.

Group BL1 BL2 V1 V2 V3 V4 V5 V6


0 18.00 18.00 17.00 18.00 15.00 17.00 14.00 15.00
0 25.11 27.00 26.00 23.00 18.00 17.00 12.00 10.00
0 19.00 16.00 17.00 14.00 -9.00 -9.00 -9.00 -9.00
0 24.00 17.00 14.00 23.00 17.00 13.00 12.00 12.00
0 19.08 15.00 12.00 10.00 8.00 4.00 5.00 5.00
0 22.00 20.00 19.00 11.54 9.00 8.00 6.82 5.05
0 28.00 16.00 13.00 13.00 9.00 7.00 8.00 7.00
0 24.00 28.00 26.00 27.00 -9.00 -9.00 -9.00 -9.00
0 27.00 28.00 26.00 24.00 19.00 13.94 11.oo 9.00
0 18.00 25.00 9.00 12.00 15.00 12.00 13.00 20.00
0 23.00 24.00 14.00 -9.00 -9.00 -9.00 -9.00 -9.00
0 21.00 16.00 19.00 13.00 14.00 23.00 15.00 11.00
0 23.00 26.00 13.00 22.00 -9.00 -9.00 -9.00 -9.00
0 21.00 21.00 7.00 13.00 -9.00 -9.00 -9.00 -9.00
0 22.00 21.00 18.00 -9.00 -9.00 -9.00 -9.00 -9.00
0 23.00 22.00 18.00 -9.00 -9.00 -9.00 -9.00 -9.00
0 26.00 26.00 19.00 13.00 22.00 12.00 18.00 13.00
0 20.00 19.00 19.00 7.00 8.00 2.00 5.00 6.00
0 20.00 22.00 20.00 15.00 20.00 17.00 15.00 13.73
0 15.00 16.00 7.00 8.00 12.00 10.00 10.00 12.00
0 22.00 21.00 19.00 18.00 16.00 13.00 16.00 15.00
0 24.00 20.00 16.00 21.00 17.00 21.00 16.00 18.00
0 -9.00 17.00 15.00 -9.00 -9.00 -9.00 -9.00 -9.00
0 24.00 22.00 20.00 21.00 17.00 14.00 14.00 10.00
0 24.00 19.00 16.00 19.00 -9.00 -9.00 -9.00 -9.00
0 22.00 21.00 7.00 4.00 4.19 4.73 3.03 3.45
0 16.00 18.00 19.00 -9.00 -9.00 -9.00 -9.00 -9.00
1 21.00 21.00 13.00 12.00 9.00 9.00 13.00 6.00
1 27.00 27.00 8.00 17.00 15.00 7.00 5.00 7.00
1 24.00 15.00 8.00 12.27 10.00 10.00 6.00 5.96
1 28.00 24.00 14.00 14.00 13.00 12.00 18.00 15.00
Sample Approaches to the Analysis of.. . 111

Table 5.1 (Continued)

Group BL1 BL2 V1 V2 V3 V4 V5 V6


1 19.00 15.00 15.00 16.00 11.00 14.00 12.00 8.00
1 17.00 17.00 9.00 5.00 3.00 6.00 0.00 2.00
1 21.00 20.00 7.00 7.00 7.00 12.00 9.00 6.00
1 18.00 18.00 8.00 1.00 1.00 2.00 0.00 1.00
1 24.00 28.00 11.oo 7.00 3.00 2.00 2.00 2.00
1 21.00 21.00 7.00 8.00 6.00 6.50 4.64 4.97
1 19.00 18.00 8.00 6.00 4.00 11.oo 7.00 6.00
1 28.00 27.46 22.00 27.00 24.00 22.00 24.00 23.00
1 23.00 19.00 14.00 12.00 15.00 12.00 9.00 6.00
1 21.00 20.00 13.00 10.00 7.00 9.00 11.oo 11.00
1 18.00 16.00 17.00 26.00 -9.00 -9.00 -9.00 -9.00
1 22.61 21.00 19.00 9.00 9.00 12.00 5.00 7.00
1 24.24 23.00 11.00 7.00 5.00 8.00 2.00 3.00
1 23.00 23.00 16.00 13.00 -9.00 -9.00 -9.00 -9.00
1 24.84 24.00 16.00 15.00 11.00 11.00 11.00 11.00
1 25.00 25.00 20.00 18.00 16.00 9.00 10.00 6.00
1 15.00 22.00 15.00 17.57 12.00 9.00 8.00 6.50
1 26.00 20.00 7.00 2.00 1.00 0.00 0.00 2.00
1 22.00 20.00 12.13 8.00 6.00 3.00 2.00 3.00
1 24.00 25.00 15.00 24.00 18.00 15.19 13.00 12.32
1 22.00 18.00 17.00 6.00 2.00 2.00 0.00 1.oo
1 27.00 26.00 1.00 18.00 10.00 13.00 12.00 10.00
1 22.00 20.00 27.00 13.00 9.00 8.00 4.00 5.00
1 20.00 17.00 20.00 10.00 8.89 8.49 7.02 6.79
1 22.00 22.00 12.00 -9.00 -9.00 -9.00 -9.00 -9.00
1 20.00 22.00 15.38 2.00 4.00 6.00 3.00 3.00
1 21.00 23.00 11.oo 9.00 10.00 8.00 7.00 4.00
1 17.00 17.00 15.00 -9.00 -9.00 -9.00 -9.00 -9.00
1 18.00 22.00 7.00 12.00 15.00 -9.00 -9.00 -9.00
1 23.00 26.00 24.00 -9.00 -9.00 -9.00 -9.00 -9.00
~~ ~~ ~ ~

0 = placebo, 1 = active, -9.00 indicates a missing values.


112 Design and Analysis of Clinical Trials

persisted for up to 18 months post-natally. Thirty-four of the women


were randomly allocated to the active treatment: 3 months of trans-
dermal 170-oestradiol 200 pg daily alone, followed by 3 months with
added cyclical dydrogeterone 10 mg daily for 12 days each month.
The remaining 27 women received placebo patches and tablets ac-
cording to the same regimen. The main outcome variable was a
composite measure of depression recorded on two occasions before
randomisation, and on six two monthly visits after randomisation.
Not all of the 61 women had the depression variable recorded on all
eight scheduled visits.
In Fig. 5.1, the data of all the 61 women are shown, separated
into treatment groups. Lines connect the repeated observations for
each woman. This simple graph makes a number of features of the
data readily apparent. First, almost all the women are becoming
less depressed. Second, the women who are most depressed at the
beginning of the study tend to be most depressed throughout. This
phenomenon is generally referred to as trucking. Third, there are
substantial individual differences and variability appears to increase
over the course of the trial.
The tracking phenomenon can be seen rather more clearly in a
plot of the standardised values of each observation, i.e., the values
obtained by subtracting the relevant visit mean from the original
observation and then dividing by the corresponding visit standard
deviation. The resulting graph appears in Fig. 5.2.
With large numbers of observations, graphical displays of individ-
ual response profiles are of little use and investigators then commonly
produce graphs showing the average profiles for each treatment group
along with some indication of the variation of the observations at
each time point. Fig. 5.3 shows such a plot for the oestrogen patch
trial example. The general decline in depression values over time in
both the active and placebo groups is apparent as is the lower level
of depression scores in the former group. An alternative t o this type
of plot is to graph side-by-side box plots of the observations at each
time point - see Fig. 5.4.
Simple Approaches t o the Analysis of.. . 113

Oestrogen patch trial individual profiles

Active

Before treatmer

?!
O J

I I I I I I I I
vl v2 v3 v4 v5 v6 v7 v8

Visit

Fig. 5.1. Individual response profiles by treatment group for the oestrogen
patch data.

An important aspect of the longitudinal data collected in many


clinical trials is the degree of association of the repeated measure-
ments. Finding a parsimonious model which describes the pattern
of associations accurately is often a crucial part of fitting the models
to be described in Chapters 6 and 7. It is frequently useful to have
some graphical display to guide the search for a realistic model; one
,
114 Design and Analysis of Clinical Trials

Oestrogen patch data


~ i s t a n d a r d i z e values
d

prel pre2 post1 post2 post3 post4 post5 post

Visit

Fig. 5.2. Individual response profiles for oestrogen patch data after stan-
dardisation.

display which can be helpful is the scatterplot matrix in which scat-


terplots of pairs of repeated measurements are placed together in a
grid. Such a plot for the data from the trial of oestrogen patches
is shown in Fig. 5.5. We see that the degree of correlation between
the pairs of observations generally decreases as the time between
them increases.
As an illustration of what can be achieved with graphical displays
on even very large sets of data, we will use an example reported by
Zeger and Katz (1998) involving an investigation of the effects of
vitamin A supplementation on Nepali pre-school children’s morbidity
and mortality. Figure 5.6 shows growth data collected in the study
from children receiving placebos. Each child was measured on up to
5 visits at four monthly intervals and the data in Fig. 5.6 consists
Simple Approaches to the Analysis o f . . . 115

, I I I I I

Prel Pre2 v1 v2 v3 v4 v5 V6
Visit(months)

Fig. 5.3. Mean response profiles for active and placebo groups in the
oestrogen patch trial.

of 11290 observations on 2 237 children. Since here connecting the


repeated measurements for all children would produce a completly
useless graphic, the measurements of a subset of 100 children have
been joined to communicate the degree of consistency across time in
a child’s weight as well as variation in weight among children. Half
of these children were selected at random and the remaining 50 were
children having average weights extreme for their age. The following
characteristics of the data are apparent from Fig. 5.6.
0 The average weight of the Nepali children increases by about
one kilogram per month for the first six months, and then the
growth rate slows dramatically to less than a quarter of the
original rate.
0 There is much greater variability in weight across children than
across time for a given child, i.e., there is strong tracking of
children’s weight so that repeated observations on the same
child will be highly correlated.
116 Design and Analysis of Clinical Trials

Boxplots for placebo

“1 Prel Prep V1
&

v2
i -

v3

v4
W

v5
L L

V6

Boxplots for active

W Y Y Y
w w

Prel Pie2 V1 v2 v3 v4 v5 V6

Fig. 5.4. Box-plots for data from the oestrogen patch data.

0 There is some indication that the rate of growth is greater for


larger children as evidenced by more positive slopes above the
average curve than below.
Sample Approaches to the Analysis o f . . . 117

16 20 2. 28 0 5 10 15 ZO 5 10 IS 20

1010 1 0 01 1 1 01 1
0 0 0 0
0

I i 0 100
a 1001
I 0 ,lo 11°
1 1
I 10 0

5 10 ar

Fig. 5.5. Scatterplot matrix for the data from the oestrogen patch trial.
118 Design and Analysis of Clinical R i a l s

Nepal Children's Growth

0 20 40 60 80
Age (months)

Fig. 5.6. Scatterplot of 11290 weights on 2 337 Nepali children. Bold dots
indicate a smoothing spline with 22 equivalent degrees of freedom as a n
estimate of the mean weigh at each age. The repeated observations for 100
children are connected - 50 chosen at random and 50 with extreme mean
weights for their age (taken with permission from Zeger and Katz, 1998).

5.3. TIME-BY-TIME ANALYSIS OF


LONGITUDINAL DATA

A time-by-time analysis of the longitudinal data consists of T sep-


arate analyses, one for each sub-set of data corresponding to each
observation time. If only two treatment groups are being compared,
each analysis consists of a t-test (or if thought necessary, its nonpara-
metric equivalent) to assess the difference in the means (medians) of
the two groups at each time point. If more than two groups are
Sample Approaches to the Analysis of.. . 119

Table 5.2. Time-by-Time Analysis of Data from Oestrogen Patch


Trial.
Group v1 v2 v3 v4 v5 V6
Placebo Mean 16.48 15.89 14.13 12.27 11.40 10.89
SD 5.28 6.12 4.97 5.85 4.44 4.68
n 27 22 17 17 17 17
Active Mean 13.37 11.74 9.13 8.83 7.31 6.59
SD 5.56 6.57 5.47 4.67 5.74 4.73
n 34 31 29 28 28 28
95% CI (0.31, (0.59, (1.77, (0.28, (0.83, (1.40,
5.91) 7.71) 8.23) 6.60) 7.34) 7.20)

involved, an analysis of variance is applied to the data at each ob-


servation time (or again perhaps some suitable nonparametric alter-
native). Relevant covariate information might be incorporated into
each analysis.
As an illustration of this approach, Table 5.2 shows the results of
applying t-tests and calculating confidence intervals for the available
data at each of the six post-randomisation visits in the oestrogen
patch trial. At each visit, there is a significant difference in the
mean depression scores of the active and placebo groups.
Although Finney (1990) suggests that the time-by-time approach
to the analysis of the longitudinal data might be quite useful if the
occasions are few and the intervals between them are large, it has,
in general, several serious drawbacks. One problem is that the struc-
ture of the data is ignored; at no stage does this type of analysis use
the information that indicates which observations are from the same
individual. Consequently, the standard errors are based on between-
subject variation only, and this is unlikely to lead to an analysis
that is fully efficient. A further problem is that inferences made
within each of the T separate analyses are not independent, nor is it
clear how they should be combined. Simply assuming that the tests
give independent information about group differences is clearly not
120 Design and Analysis of Clinical Trials

sensible, as is demonstrated by considering what would happen if the


repeated measurements were made more frequently. The number of
significance tests performed would rise accordingly, but the increase
in information about the difference between treatments is likely to
be small.
Repeated testing also implicitly assumes that each time point is of
interest in its own right; this is unlikely to be so in most applications,
where the real interest will be in something more global. It is highly
questionable whether the hypotheses being tested by the time-by-
time approach, namely the equality of the group means at each time
point, are of interest. Such an analysis does not give an overall
answer to whether or not there is a treatment difference and provides
no single estimate of the treatment effect.
A time-by-time analysis is not, in general, appropriate for analy-
sing data collected over time in a clinical trial, since it simply ignores
the longitudinal nature of the observations. There may, however,
be some instances where the method can lead to useful results if it
concentrates on exactly the right feature of the data, as shown in a
non clinical trial context by Kenward (1987). Trying to find some
feature of the data that characterises concisely their structure is also
central to another simple, but in this case often very useful, method
for the analysis of longitudinal data from clinical trials, namely the
response feature or summary measures approach.

5.4. RESPONSE FEATURE ANALYSIS OF


LONGITUDINAL DATA

The use of summary measures is one of the most important and


straightforward approaches to the analysis of longitudinal data (at
least when there are few missing values, and when the repeated mea-
surements on each subject are made at the same time points). If
the measurements made on the ith individual in the study are writ-
ten as the vector x!,= [ z i l , .. . ,z~T],
then a scalar-valued function f
is chosen so that si = f(xi) captures some essential feature of the
Simple Approaches to the Analysis of.. . 121

response over time (see Section 5.3.1 for more details). In this way,
the essentially multivariate nature of the repeated observations is re-
duced to a univariate one. This approach has been in use for many
years, and is described in Oldham (1962), Yates (1982) and Matthews
et al. (1990). Various aspects of response feature analysis will be con-
sidered in this section including how to incorporate covariates, how
to deal with unbalanced data, and the implication for the method of
having missing values. We begin, however, with perhaps the most
important consideration, namely the choice of summary measure.

5.4.1. Choosing Summary Measures


To begin with, we shall assume that the response variable observed
over time is continuous (or quasi-continuous) but not necessarily
normally distributed. The key step to a successful response fea-
ture analysis is the choice of a relevant summary measure f. In most
cases, the decision over what is a suitable measure should to be made
before the data are collected. The chosen summary measure needs to
be relevant to the particular questions of interest in the study and in
the broader scientific context in which the study takes place. In some
longitudinal studies, more than a single summary measure might be
deemed relevant or necessary, in which case the problem of combined
inference may need t o be addressed. More often in practice, however,
it is likely that the different measures will deal with substantially
different questions so that each will have a natural interpretation in
its own right.
A wide range of possible summary measures have been proposed.
Those given in Table 5.3, for example, were suggested by Matthews
et al. (1990). Frison and Pocock (1992) argue that the average re-
sponse to treatment over time is often likely to be the most relevant
summary statistic in treatment trials. In some cases the response
on a particular visit may be chosen as the summary statistic of most
interest, but this must be distinguished from the generally flawed ap-
proach which separately analyses the observations at each and every
time point.
122 Design and Analysis of Clinical R a l s

Table 5.3. Possible Summary Measures (from Matthews et al.,


1990).

Type of Data Questions of Interest Summary Measure


Peaked Is overall value of outcome Overall mean (equal
variable the same in time intervals) or area
different groups? under curve (unequal
intervals)
Peaked Is maximum (minimum) Maximum (minimum)
response different value
between groups?
Peaked Is time to maximum Time to maximum
(minimum) response (minimum) response
different between groups?
Growth Is rate of change of outcome Regression coefficient
different between groups?
Growth Is eventual value of outcome Final value of outcome
different between groups? or difference between
last and first values
or percentage change
between first and
last values
Growth Is response in one group Time to reach a
delayed relative to the particular value (e.g.
other? a fixed percentage of
baseline)

As our first example of the application of the summary measure


technique, we will use the data shown in Table 5.4. These data arise
from a n investigation of the use of lecithin, a precursor of choline,
in the treatment of Alzheimer’s disease. Traditionally, it has been
assumed that this condition involves a n inevitable and progressive
deterioration in all aspects of intellect, self-care, and personality.
Simple Approaches to the Analysis of.. . 123

Table 5.4. Data from Trial of Lecithin for the Treat-


ment of Alzheimer’s Disease.
Group v1 v2 v3 v4 v5
1 20 19 20 20 18
1 14 15 16 9 6
1 7 5 8 8 5
1 6 10 9 10 10
1 9 7 9 5 6
1 9 10 9 11 11
1 7 3 7 6 3
1 18 20 20 23 21
1 6 10 10 13 14
1 10 15 15 15 14
1 5 9 7 3 12
1 11 11 8 10 9
1 10 2 9 3 2
1 17 12 14 15 13
1 16 15 13 7 9
1 7 10 4 10 5
1 5 0 5 0 0
1 16 7 7 6 10
1 5 6 9 5 6
1 2 1 1 2 2
1 7 11 7 5 11
1 9 16 17 10 6
1 2 5 6 7 6
1 7 3 5 5 5
1 19 13 19 17 17
1 7 5 8 8 6
2 9 11 14 11 14
2 6 7 9 12 16
2 13 18 14 20 14
2 9 10 9 8 7
2 6 7 4 5 4
124 Design and Analysis of Clinical Trials

Table 5.4 (Continued)


Group V1 v2 v3 v4 v5
2 11 11 5 10 12
2 7 10 11 8 5
2 8 18 19 15 14
2 3 3 3 1 3
2 4 10 9 17 10
2 11 10 5 15 16
2 1 3 2 2 5
2 6 7 7 6 7
2 0 3 2 0 0
2 18 19 15 17 20
2 15 15 15 14 12
2 14 11 8 10 8
2 6 6 5 5 8
2 10 10 6 10 9
2 4 6 6 4 2
2 4 13 9 8 7
2 14 7 8 10 6
1 = placebo, 2 = lecithin

Recent work suggests that the disease involves pathological changes


in the central cholinergic system, which it might be possible t o rem-
edy by long-term dietary enrichment with lecithin. In particular,
the treatment might slow down or perhaps even halt the memory
impairment usually associated with the condition. Patients suffering
from Alzheimer’s disease were randomly allocated to receive either
lecithin or placebo for a six-month period. A cognitive test score
giving the number of words recalled from a previously studied list
was recorded at the start, at one month, at two months, at four
months and at six months. As a summary measure for these data,
we shall use the maximum number of words recalled on any of the five
occasions of testing. (The ‘maximum’ was chosen on clinical advice,
Sample Approaches to the Analysis of.. . 125

Table 5.5. Results from Response Feature


Approach on Data from Lecithin Dial Using
Maximum Number of Words Over the Five
Visits as Summary Measure.
Placebo Active
Mean 11.9 11.8
SD 5.0 5.2
n 26 22
95% confidence interval for difference - (-2.9,3.1)

although it may have less desirable statistical properties, for example,


its variance will be greater, than the mean.) The results of applying a
t-test to the summary measures of each patient in the two treatment
groups is given in Table 5.5. There is no evidence of any treatment
effect.
As a further illustration of the response feature approach, we
shall again use the data from the oestrogen patch trial given in
Table 5.1. The chosen summary measure here is the mean of the
post-randomisation measures. This immediately gives rise to a com-
plication since not all the women in the trial have observations on
all six occasions. The summary measure approach is often suffi-
ciently flexible to accomodate missing values relatively simply; here,
for example, we could take the mean of the available observations
for each patient. One alternative would be to use only the women
with observations on all six post-randomisation occasions and an-
other possibility would be to use the last recorded value for a woman
for all the missing values, the so-called last obsermation carried for-
ward (LOCF) approach. (Missing values might also be imputed by,
for example, substituting relevant mean values of using some more
complex procedure.) The results obtained from using each approach
are shown in Table 5.6. In this case, the three confidence intervals
for the treatment difference are very similar. In most cases when
using the response feature approach, however, particularly when
the proportion of missing observations is small, using the available
126 Design and Analysis of Clinical l?raals

Table 5.6. Results of Using Summary Mea-


sure Procedure with Mean Depression Score
as Summary Measure in Oestrogen Patch Data
with Various Approaches to Dealing with the
Missing Values.
(1) Mean of usable values
Placebo Active
mean 14.76 10.55
sd 4.56 5.36
n 27 34
95% confidence interval for difference - (1.61,6.79)
(2) Complete cases
Placebo Active
mean 13.38 9.30
sd 4.28 4.57
n 17 28
95% confidence interval for difference - (1.33,6.82)
(3) Last observation carried forward
Placebo Active
mean 14.95 10.66
sd 4.66 5.56
n 27 34
95% confidence interval for difference - (1.62,6.96)

observations to calculate the summary measure is recommended al-


though with accompanying caveats about two potential problems:

0 If the summary measure si are based on observations xi made


at considerably different sets of time points (either by design
or because of the occurrence of missing values), then the stan-
dard t-test or analysis of variance assumption of a common
Sample Approaches to the Analysis of.. . 127

variance for all observations can no longer be true. Means


based on different numbers of observations, regression slopes
based on differently located observations, and maxima within
sets of observations of different sizes, will not share common
distributions. The likely importance and impact of this prob-
lem will need t o be judged in each particular application.
(Matthews, 1993, describes a refinement of the response fea-
ture approach which goes some way to dealing with the prob-
lem of observations at irregular time points. The central idea
of Matthew’s proposal is to introduce some form of weighting
into the estimation of the treatment difference.)
0 The type of missing value (see Chapter 2) has implications for
the suitability of the summary measure approach. When the
observations are missing completely at random, calculating
the chosen summary measure from the available observations
is a valid and acceptable procedure as is using only complete
cases, although the latter will be less efficient particularly if
there, is a considerable proportion of intended observations
that are missing. But if the missing values are thought to be
other than MCAR, response feature analysis may be seriously
misleading. If, for example, interest focuses on the maximum
depression score over time, and some adverse effect of severe
depression stops them from being observed (the patient cannot
attend the clinic perhaps), then simply using the maximum
of the observations that were obtained would clearly not be
accept able.
Imputing missing values before calculating summary measures is
also not without its problems. Using the last observation carried
forward (LOCF) method in which each missing value is substituted
by the subject’s last available assessment of the same type is often
used, particularly in the pharmaceutical industry. But the method
appears to have little in its favour except that it records what has
been achievable with a particular patient. The procedure involves,
however, highly unlikely assumptions, for example, that the expected
128 Design and Analysis of Clinical Trials

value of the (unobserved) post dropout responses remain at their last


recorded value.
Imputation methods need to be carefully chosen to avoid biased
estimates from filled-in data. Also, imputation invents data, and
analysing filled-in data as if they were complete leads to overstate-
ment of precision, i.e., standard errors are underestimated, stated
p-values of tests are too small, and confidence intervals do not cover
the true parameter at the stated rate. More will be said about im-
putation in other contexts in Chapters 6 and 7.
Although the simple summary measures listed in Table 5.3 are
the ones most likely to be commonly used, Gornbein et al. (1992)
and Diggle et al. (1994) urge a more imaginative use of the summary
measure approach, with the fitting of scientifically interpretable non-
linear models to the observations collected on each individual and the
use of derived parameter stimates as the summary measures for anal-
ysis. An example given by Gornbein et al. (1992) involves a study to
compare the relationship between parathyroid hormone (PTH) and
serum ionized calcium (ICa) in normal teenagers versus teenagers
undergoing dialysis for end stage renal disease. As ICa levels are
artificially lowered or raised in the serum, the parathyroid gland
senses the change and responds by altering the production of PTH.
Increased levels of PTH cause more calcium to be released from the
bones. Decreased levels result in more calcium absorption by the
bones. In this way, the body attempts to restore the serum ICa to
a ‘setpoint’ level. Based on this biological model, the investigators
expected to see a sigmoid relationship between ICa and PTH.
The data were incomplete for some subjects as the PTH sam-
ples taken at a given ICa level were sometimes contaminated, lost
or spoiled. In addition, the same ICa levels were not used for all
subjects. The following normal sigmoid model was fitted to each
partcipant’s repeated observations:

Y = PO+ pl/[1 + ( z / P ~ ) ~ ~ P ( P+~ ) ]


€, - N ( O 0, 2 ) (5.1)
Simple Approaches to the Analysis of.. . 129

100 m
I I

1 .o 1.1 1.2 1.3 1.4 1 .s 1.6

C a + + Irnmol/Ll

Fig. 5.7. Observations and fitted curve for one subject’s data in study of
parathyroid hormone (taken with permission from Gornbein et al., 1992).

(The observation and fitted curve for one subject are shown in
Fig. 5.7.)
Each subject’s original data are now summarised by the four pa-
rameters po,/31,pz and ps. Normal volunteers can be compared to
diseased subjects by comparing these four parameters, rather than
by comparing mean profiles. (Since there are four parameters, a
multivariate test such as Hotelling’s T2might be needed.) As these
parameters have a physical meaning, this sort of comparison is gen-
erally easier to interpret than one involving mean profiles.

5.4.2. Incorporating Pre-Randomisat ion,


Baseline Measures into the Response
Feature Approach
Baseline measurements of the outcome variable, if available, can be
used in association with the response feature method in a number
of ways. Frisson and Pocock (1992), for example, suggest three
130 Design and Analysis of Clinical Trials

possibilities when the average response over time is the chosen sum-
mary measure:

0 POST - an analysis that ignores the pre-randomisation values


available and analyses only the mean of each subject’s post-
randomisat ion responses;
0 CHANGE - an analysis that uses the differences between
the means of each subject’s post- and pre-randomisation re-
sponses;
0 ANCOVA - here between subject variation in baseline mea-
surements is taken into account by using the mean of the
baseline values for each subject as a covariate in a linear model
for the comparison of post-randomisation means.

The mathematical details of each approach in the simplest case


of a single baseline measure and a single post-randomisation measure
are shown in Table 5.7. It is clear that the estimation of the treatment
effect in each case differ in how the observed difference at outcome
is adjusted for the baseline difference. POST, for example, relies on
the fact that in a randomised trial, in the absence of a difference
between treatments, the expected value of the mean difference at
outcome is zero. Hence, the factor by which the observed outcome
needs correcting in order to judge the treatment effect, is also zero.
The CHANGE approach corresponds to the assumption that the
difference at outcome, in the absence of a treatment effect, is ex-
pected to be equal to the difference in the means of the baselines. In
the context of many clinical trials, however, this assumption may be
false as is well documented by a variety of authors, e.g., Chuang-Stein
and Tong (1997), Chuang-Stein (1993) and Senn (1994a, 1994b). The
difficulty primarily involves the regression to the mean, phenomenon.
This refers to the process that occurs as transient components of an
initial score are dissipated over time. Selection of high scoring indi-
viduals for entry into a trial necessarily also selects for individuals
with high values of any transient component that might contribute
to that score. Remeasurement during the trial will tend to show
Simple Approaches to the Analysis of.. . 131

Table 5.7. POST, CHANGE and ANCOVA compared.

We will consider a randomised parallel groups clinical trial for which baseline
and final values are available for each member of the control and treatment
groups.
* Measurements on members of the treatment group are represented by X t ,
(baselines) and Yt, (outcomes), i = 1,.. . ,nt. Similarly, measurements in the
control group are represented by Xci and Yci,i = 1,.. . ,nc.
Suppose that the covariance matrix, X X , Yin , the two groups is identical and
given by:

When POST
- -
is the method of analysis, the effect of treatment is estimated as
PraW= Yt - Y, with variance var (iraw) = qu? where q = 2- +
nt 1
n, .
When CHANGE is used, the effect of treatment is estimated as ?-change =
(s - Xt)- (pc- X c ) = (c - pc) - ( X t - X,) with variance var(FCjchange)=
q(& + ff -; 2paxay).
A more general estimator of the treatment effect is given by Fo = (pt - Yc)-
P(xt- X,) which has variance var(+p) = q[(Oux- p u ~ +) (1~ - p 2 ) u $ ] . AN-
COVA corresponds to choosing a suitable value for p, namely /?’= p a y l u x .
The raw-outcomes and change-score estimator are special forms of the general
estimator with 0 = 0 and ,B = 1.
The three methods POST, CHANGE and ANCOVA can now be seen merely
as ways of adjusting the observed difference at outcome using the baseline
difference.
POST relies on the fact that in a randomised trial, in the absence of a treat-
ment effect, the expected value of the mean difference at outcome is zero.
Hence the factor by which the observed outcome needs correcting, in order to
judge the treatment effect, is also zero.
CHANGE corresponds to the assumption that the difference at outcome, in
the absence of a treatment effect, is expected to be the diffference in the
means of the baselines. This assumption is, in fact, false (see, for example,
Senn, 1994).
ANCOVA allows for a more general system of predicting what the outcome
difference would have been in the absence of any treatment effect as a function
of the mean difference at baseline

(The above is an abbreviated version of Senn, 1998, from where a more detailed
account is available.)
132 Design and Analysis of Clinical Dials

a declining mean value for such groups. Consequently, groups that


initially differ through the existence of transient phenomena such as
some forms of measurement error, will show a tendency to have con-
verged on remeasurement. Randomisation ensures only that treat-
ment groups are similar only in terms of expected values and so
may actually differ not just in transient phenomena but also in more
permanent components of the observed scores. Thus, while the dis-
sipation of transient components may bring about regression to the
mean phenomena such as those previously described, the extent of
regression and the mean value to which the separate groups are re-
gressing need not be expected to be the same. Analysis of covariance
(ANCOVA) allows for such differences. The use of ANCOVA allows
for some more general system of predicting what the outcome differ-
ence would have been in the absence of any treatment effect, as a
function of the mean difference at baseline.
Frison and Pocock (1992) compare the three approaches when
different numbers of pre-randomisation measurements are available.
With a single baseline measure of the outcome variable, they show
that analysis of covariance is more powerful than both analysis of
change scores and analysis of post-randomisation means only, except
when the correlations between the repeated measurements are small.
Using the mean of several pre-randomisation measures (if available),
makes the analysis of covariance even more efficient if there are sub-
stantial correlations between the repeated observations.
The differences between the three approaches can be illustrated
by comparing power curves calculated using the results given in Fri-
son and Pocock (1992). Figures 5.8, 5.9 and 5.10 show some examples
for the situation with two treatment groups, two pre-randomisation
values and six post-randomisation observations and varying degrees
of association between the repeated observations (the correlations
between pairs of repeated measurements are assumed equal in the
calculation of these power curves). From these curves, it can be seen
that the sample size needed to achieve a particular power for detect-
ing a standardised treatment difference of 0.5 is always lower with
Simple Approaches to the Analysis of.. . 133

Power curves for three methods


of analysing repeated measure designs

,,..----/---
__--
,. *.-
_.=

I- Ancova
Post scores
___.-
Change scores

Npost=6
Std. diff.=0.5
RhOS.3
Alpha=O.OB

I I I I
0 10 20 30 40 50

Sample size

Fig. 5.8. Power curves for POST, CHANGE and ANCOVA.

analysis of covariance, and in some cases, substantially lower. As the


correlation between the repeated observations increases, CHANGE
approaches ANCOVA in power, with both being considerably better
than POST. With a low correlation of 0.2, however, CHANGE does
less well than simply dealing with post-randomisat ion values only.
(When the correlation is zero, ANCOVA and POST are essentially
equivalent. )
134 Design and Analysis of Clinical Dials

Power curves for three methods


of analysing repeated measure designs

0 20 40 60

Sample size

Fig. 5.9. Power curves for POST, CHANGE and ANCOVA.

The results of applying each of POST, CHANGE and ANCOVA


to the oestrogen patch example, using the mean of available post-
randomisation observation as the response and the mean of avail-
able pre-randomisation measures as baseline, are shown in Table 5.8.
Here, all three approaches indicate a substantial treatment effect.

5.4.3. Response Feature Analysis when the


Response Variable is Binary
Table 5.9 shows the data collected in a clinical trial comparing two
treatments for a respiratory illness (Davis, 1991). In each of the two
Simple Approaches t o the Analysis of.. . 135

Power curves for three methods


of analysing repeated measure designs

_..----

,/’

,/”

,.//
/

/
/ i’
..... Post scores
Change scores
i
,
I i

// Npre=2
/ Npk6
Std. diff.=0.5
Rhod.9
Alpha=O.O5

I I I I I
0 20 40 60 80

Sample size

Fig. 5.10. Power curves for POST, CHANGE and ANCOVA.

centres, eligible patients were randomly assigned to active treatment


or placebo. During treatment, the respiratory status (categorised
as 0 = poor,1 = good) was determined at four visits. There were
111 patients (54 active, 57 placebo) with no missing data for re-
sponses or covariates. Some detailed analyses of these data will be
described in Chapter 7, but here we shall consider how the response
feature approach could be used. We might, of course, simply ig-
nore the binary nature of the response variable and compare the
‘mean’ responses over time in the two treatment groups by a t-test.
136 Design and Analysis of Clinical Trials

Table 5.8. Application of POST, CHANGE


and ANCOVA to Mean Depression Scores
used as Response and Covariate (means of us-
able values).
(1) POST
95% confidence interval for difference - (1.61,6.79)
(Details given in Table 5.6).
(2) CHANGE
Placebo Active
mean -6.51 -11.07
sd 4.38 5.49
n 27 34
95% confidence interval for difference - (1.97,7.16)
(3) ANCOVA
Placebo Active
Adjusted Mean Adjusted Mean
14.85 10.47
Mean square from analysis of covariance is 23.29
based on 58 degrees of freedom.
95% confidence interval for difference - (1.89,6.87)

Since the mean in this case is the proportion ( p ) of visits at which a


patient’s respiratory status was good, we could consider performing
the test after taking some appropriate transformation, for example,
arcsin(p) or arcsin(@). All such tests indicate that there is a sub-
stantial difference between the two treatments. A linear regression
of the arcsin transformed proportion of positive responses over the
four post-baseline measurement occasions might be used to assess
the effects of the baseline measurement, age, sex and centre. The
results are shown in Table 5.10
A more satisfactory analysis can be achieved by using the GLM
approach described in the previous chapter. A standard logistic
Simple Approaches to the Analysis o f . . . 137

Table 5.9. Respiratory Disorder Data.


~~

Patient Treat. Sex Age BL V1 V2 V3 V4


Centre 1
1 P M 46 0 0 0 0 0
2 P M 28 0 0 0 0 0
3 A M 23 1 1 1 1 1
4 P M 44 1 1 1 1 0
5 P F 13 1 1 1 1 1
6 A M 34 0 0 0 0 0
7 P M 43 0 1 0 1 1
8 A M 28 0 0 0 0 0
9 A M 31 1 1 1 1 1
10 P M 37 1 0 1 1 0
11 A M 30 1 1 1 1 1
12 A M 14 0 1 1 1 0
13 P M 23 1 1 0 0 0
14 P M 30 0 0 0 0 0
15 P M 20 1 1 1 1 1
16 A M 22 0 0 0 0 1
17 P M 25 0 0 0 0 0
18 A F 47 0 0 1 1 1
19 P F 31 0 0 0 0 0
20 A M 20 1 1 0 1 0
21 A M 26 0 1 0 1 0
22 A M 46 1 1 1 1 1
23 A M 32 1 1 1 1 1
24 A M 48 0 1 0 0 0
25 P F 35 0 0 0 0 0
26 A M 26 0 0 0 0 0
27 P M 23 1 1 0 1 1
28 P F 36 0 1 1 0 0
29 P M 19 0 1 1 0 0
30 A M 28 0 0 0 0 0
31 P M 37 0 0 0 0 0
138 Design and Analysis of Clinical Trials

Table 5.9 (Continued)


Patient Treat. Sex Age BL V1 V2 V3 V4
32 A M 23 0 1 1 1 1
33 A M 30 1 1 1 1 0
34 P M 15 0 0 1 1 0
35 A M 26 0 0 0 1 0
36 P F 45 0 0 0 0 0
37 A M 31 0 0 1 0 0
38 A M 50 0 0 0 0 0
39 P M 28 0 0 0 0 0
40 P M 26 0 0 0 0 0
41 P M 14 0 0 0 0 1
42 A M 31 0 0 1 0 0
43 P M 13 1 1 1 1 1
44 P M 27 0 0 0 0 0
45 P M 26 0 1 0 1 1
46 P M 49 0 0 0 0 0
47 P M 63 0 0 0 0 0
48 A M 57 1 1 1 1 1
49 P M 27 1 1 1 1 1
50 A M 22 0 0 1 1 1
51 A M 15 0 0 1 1 1
52 P M 43 0 0 0 1 0
53 A F 32 0 0 0 1 0
54 A M 11 1 1 1 1 0
55 P M 24 1 1 1 1 1
56 A M 25 0 1 1 0 1
Centre 2
1 P F 39 0 0 0 0 0
2 A M 25 0 0 1 1 1
3 A M 58 1 1 1 1 1
4 P F 51 1 1 0 1 1
5 P F 32 1 0 0 1 1
6 P M 45 1 1 0 0 0
Simple Approaches to the Analysis of.. . 139

Table 5.9 (Continued)

Patient Treat. Sex Age BL V1 V2 V3 V4


7 P F 44 1 1 1 1 1
8 P F 48 0 0 0 0 0
9 A M 26 0 1 1 1 1
10 A M 14 0 1 1 1 1
11 P F 48 0 0 0 0 0
12 A M 13 1 1 1 1 1
13 P M 20 0 1 1 1 1
14 A M 37 1 1 0 0 1
15 A M 25 1 1 1 1 1
16 A M 20 0 0 0 0 0
17 P F 58 0 1 0 0 0
18 P M 38 1 1 0 0 0
19 A M 55 1 1 1 1 1
20 A M 24 1 1 1 1 1
21 P F 36 1 1 0 0 1
22 P M 36 0 1 1 1 1
23 A F 60 1 1 1 1 1
24 P M 15 1 0 0 1 1
25 A M 25 1 1 1 1 0
26 A M 35 1 1 1 1 1
27 A M 19 1 1 0 1 1
28 P F 31 1 1 1 1 1
29 A M 21 1 1 1 1 1
30 A F 37 0 1 1 1 1
31 P M 52 0 1 1 1 1
32 A M 55 0 0 1 1 0
33 P M 19 1 0 0 1 1
34 P M 20 1 0 1 1 1
35 P M 42 1 0 0 0 0
36 A M 41 1 1 1 1 1
37 A M 52 0 0 0 0 0
38 P F 47 0 1 1 0 1
140 Design and Analysis of Clinical Trials

Table 5.9 (Continued)


Patient Treat. Sex Age BL V1 V2 V3 V4
39 P M 11 1 1 1 1 1
40 P M 14 0 0 0 1 0
41 P M 15 1 1 1 1 1
42 P M 66 1 1 1 1 1
43 A M 34 0 1 1 0 1
44 P M 43 0 0 0 0 0
45 P M 33 1 1 1 0 0
46 P M 48 1 1 0 0 0
47 A M 20 0 1 1 1 1
48 P F 39 1 0 1 0 0
49 A M 28 0 1 0 0 0
50 P F 38 0 0 0 0 0
51 A M 43 1 1 1 1 0
52 A F 39 0 1 1 1 1
53 A M 68 0 1 1 1 1
54 A F 63 1 1 1 1 1
55 A M 31 1 1 1 1 1
Treatment: P = placebo, A = active,
Gender: M = male, F = female,
Response: 0 = respiratory status poor, 1= respiratory status good.

Table 5.10. Results from a Linear Regression of


the Arcsin Transformed Response of Positive Re-
sponses over the Four Post-baseline Measurement
Occasions for the Respiratory Data in Table 5.9.

Covariate Estimate SE P
Treatment 0.377 0.101 < 0.001
Baseline 0.580 0.103 < 0.001
Age1100 -0.402 0.386 0.3
Sex 0.039 0.133 0.8
Centre 0.205 0.107 0.06
Sample Approaches to the Analysis of.. . 141

Table 5.11. Results from a GLM for a Logit Model


Allowing for Overdispersion for the Respiratory
Data in Table 5.9.

Covariate Estimate Odds Ratio SE P


Treatment 3.544 1.208 < 0.001
Baseline 6.333 2.197 < 0.001
Age1100 0.153 0.196 0.1
Sex 1.147 0.488 0.7
Centre 1.915 0.661 0.06

regression model might be applied but since the number of occasions


on which infection was present out of the 4 visits made by each
participant is unlikely to be binomially distributed (the observations
are likely to be correlated rather than independent), we need to allow
for possible overdispersion. This is relatively straightforward using
the quasilikelihood procedure of Wedderburn (1974) to fit a GLM
for overdispersed binomial data with variance function 4p(1 - p ) / 4 ,
where q!~is estimated as described in the previous chapter.
Fitting a model with logistic link and with treatment, sex, age
and baseline respiratory status as the main effects gives the results
shown in Table 5.11. The estimated value of the scale parameter,
2.10, is substantially above one confirming the presence of overdis-
persion. The estimated odds ratio for the effect of treatment is 3.54,
with a 95% confidence interval of (1.82,6.91).
The p-values from the linear regression in Table 5.10 and the
logistic regression in Table 5.11 are comparable, but the estimates
from the logit model are on a more natural scale. There is no rela-
tively straightforward interpretation of the estimates from the model
fitted to the arcsin transformed responses, although calculation of
an odds-ratio at the mean value of the sample covariates could be
attempted.
142 Design and Analysis of Clinical Trials

5.5. SUMMARY
The methods described in this chapter provide for the exploration
and simple analysis of longitudinal data collected in the course of
a clinical trial. The graphical methods can provide insights into
both potentially interesting patterns of response over time and the
structure of any treatment differences. In addition, they can indi-
cate possible outlying observations that may need special attention.
The response feature approach to analysis has the distinct advantage
that it is straightforward, can be tailored to consider aspects of the
data thought to be particularly relevant, and produces results which
are relatively simple to understand. The method can accomodate
data containing missing values without difficulty, although it might
be misleading if the observations are anything other than missing
completely at random.
CHAPTER6

Multivariate Normal
Regression Models for
Longitudinal Data from
Clinical Trials

6.1. INTRODUCTION

It cannot be overemphasised that statistical analyses of clinical trials


should be no more complex than necessary. So even when repeated
measures data have been collected it is not always essential to a p
ply a formal repeated measures analysis. We have already described
analyses of individual summary measures such as individual mean
response scores and proportions that may often not only be statisti-
cally and scientifically adequate for the testing of simple treatment
differences, but can also be more persuasive and easier to communi-
cate than some of the more ambitious analyses that we now turn to.
Nonetheless, a typical trial does not take place in a scientific vacuum
in which a simple treatment difference is the only question of inter-
est. It is more usual that, in addition to being used as the basis of
formal evidence for efficacy or equivalence, even a phase I11 trial will
gather data that may be informative as to the mode of action of the
treatment, dose-response relationship, response heterogeneity, side
effects and so on. In addition, attrition and variation in compliance

143
144 Design and Analysis of Clinical Trials

may need to be more carefully examined than can be done within,


for example, a simple summary statistics approach.
In this chapter, models suitable for analysing longitudinal data
from clinical trials where the response variable can be assumed to
have a conditionally normal distribution will be described. Models
suitable for non normal responses will be the subject of Chapter 7.

6.2. SOME GENERAL COMMENTS


ABOUT MODELS FOR
LONGITUDINAL DATA

Diggle (1988) lists a number of desirable features for a general


method to analyse data from studies in which outcome variables are
measured at several time points. These include the following:

0 the specification of the mean response profile needs to be


sufficiently flexible to reflect both time trends within each
treatment group and differences in these time trends between
treatments. Examples of the type of mean response profiles
that may need to be modelled are shown in Fig. 6.1;
0 the specification of the pattern of correlations or covariances
of the repeated measurements needs to be flexible, but
economical;
0 the method of analysis should accommodate virtually arbi-
trary patterns of irregularly spaced time sequences within
individuals.

Although in many applications the parameters defining the


covariance structure of the observations will not be of direct in-
terest (they are often regarded as so-called nuisance parameters),
Diggle (1988) suggests that overparametrisation will lead to ineffi-
cient estimation and potentially poor assessment of standard errors
for estimates of the mean response profiles, whereas too restrictive
a specification may also invalidate inferences about the mean re-
sponse profiles when the assumed covariance structure does not hold.
Multivariate Normal Regression Models for . 145

0
Lo . In
o

-3
0

Fl
0 . m
o

E
i=
0
N - 0
N

m
- 0 a,
2 a
2a
0 - 0

0s OP OE 02 01. 0 0s 09 OE 02 oc 0

aSUOdS@J UEaw asuodsa ueayy

0
Lo I
n
0 a
2
rn
0
*
0 ' S a
Lu
0

m
0 . mo

E E
F
i=
N
0
2
0
m
z -!
(D

bb
0 !z
0s OP OE 02 01. 0 0s 09 OE OZ OL 0

asuodsa) u e a ~ asuodsai ueayy


146 Design and Analysis of Clinical Trials

In many examples, the model for the covariances and variances of


the repeated measures will need to allow for non-stationarity, with
changes (most often increases) in variances across time being partic-
ularly common. The effect of misspecifying the covariance structure
when analysing longitudinal data is investigated in Rabe-Hesketh
and Everitt, 1999.
In this chapter, we shall consider the likelihood methods for
estimating parameters and their standard errors in very general
regression-type models for longitudinal data.
Some of the advantages of using a likelihood approach are as
follows:
0 maximum likelihood estimates are principled in that they have
known statistical properties (consistency, large sample effi-
ciency) under the assumed model, which can be clearly speci-
fied and subjected to model criticism;
0 maximum likelihood estimation does not require a rectangu-
lar data matrix and hence deals directly with the problem of
unbalanced data;
0 estimates are asymptotically efficient under the assumed
model;
0 standard errors of parameter estimates based on the observed
and/or expected information matrix are available and these
automatically take into account the fact that the data are
incomplete.
Two disadvantages of the likelihood approach are:
0 maximum likelihood estimation requires the specification of a
full statistical model for the data and results may be vulnera-
ble to departures from model assumptions (though use of the
heteroscedastic consistent ‘sandwich’estimator of the parame-
ter variances introduced in Chapter 4 provides some scope for
relaxing assumptions) ;
0 maximum likelihood inferences are based on large sample the-
ory and hence may be unsuitable for small data sets.
Multivariate Normal Regression Models for . . . 147

The details of maximum likelihood estimation of the parameters


in regression models for longitudinal data will be described in Sec-
tion 6.4. Before this, however, the implications for this approach of
missing values due to patients dropping-out of the study need to be
considered.

6.3. MISSING VALUES A N D


LIKELIHOOD-BASED INFERENCE
FOR LONGITUDINAL DATA
In Chapter 3 a taxonomy of missing values occurring in longitudi-
nal data was introduced. The essential distinctions drawn between
missing values were:

MCAR ~ missing completely at random, the drop-out process


is independent of both the observed and missing values;
MAR - missing at random the drop-out process is indepen-
dent of the missing values;
0 Informative - the drop-out process is dependent on the miss-
ing values.

For likelihood-based inference, the critical distinction is between


MCAR and MAR (often referred to collectively as ignorable) and in-
formative dropout since in the former case the log-likelihood function
is separable into two terms, one involving the missing-data mecha-
nism given the observed values and one only involving the observed
values. The first of these contains no information about the distribu-
tion of observed values and can therefore be ignored for the purpose
of making inferences about these values. (Mathematical details are
given in Table 6.1.) It is important to make clear, however, that
it is the missing-data mechanism that is ignorable, n o t the individ-
uals with missing values. This is not a trivial point because one
of the most common methods of analysing unbalanced longitudinal
data remains using either analysis of variance or multivariate anal-
ysis of variance (see comments in Section 6.6) on complete cases.
148 Design and Analysis of Clinical Trials

Table 6.1. Separability of Likelihood for Non-informative Missing


Values.
Let the complete set of measurements Y * be partioned into Y * =
[Y("),Y'")],with Y(")representing the observed measurements and
Y(")the missing measurements.
Let R represent a set of indicator random variables denoting which ele-
ments of Y * are observed and which are missing.
Let f(y("),
y ( m ) r)
, represent the joint probability density function of
( Y ( " Y("),R).
),
0 This density function can be written as:

For a likelihood-based analysis, we need the joint pdf of the observable


random variables (Y'"), R), which is obtained by integrating the above
to give:
f(y(o),r) = / f(y(O),y(m))f(rly(o),
y("))dy(")

If the missing value mechanism is non informative, f(r(y("),y("))


does
not depend on y(") and the above becomes:

Taking logs of this expression gives:

L can be maximised by separate maximisation of the two terms on the


right hand side. Since the first term contains no information about the
distribution of Y ( " )we
, can ignore it for the purposes of making inferences
about Y ( " ) .

(The account above is taken from Diggle et al, 1994.)


Multivariate Normal Regression Models for . . . 149

At best (when the missing values are MCAR) such an approach is


inefficient since it leaves out observations which could legitimately
be used in the analysis; at worst (when the observations are MAR
or informative) it may lead to erroneous inferences. The likelihood
method to be described in the next section depends only on the
weaker assumption that the missing values are MAR. (The difficult
problems of distinguishing between random and informative missing
values and of analysing data containing the latter will be dealt with
in Section 6.7.)

6.4. MAXIMUM LIKELIHOOD


ESTIMATION ASSUMING
NORMALITY

Following Schluchter (1988), the essence of the likelihood approach


for longitudinal data possibly containing missing values is as follows:
Let y,’ denote the hypothetical T x 1 complete-data vector for
subject i. Let y: = [yil, . . . , yin,] be the ni x 1 vector of measure-
ments actually present for subject i , i = 1 , 2 , . . . ,n. (We assume that
measurements for subject i are made at times t i l , t i 2 , . . . ,tini,i.e., we
do not assume a common set of times for all subjects.) The yij are
assumed to be realisations of random variables y Z j which follow the
regression model;
Yz = xzp + € 2 (6.1)

for i = 1,. . . ,n, where Xi is a ni x p design matrix, p is a p x 1 vector


of unknown regression coefficients. We assume that the conditional
distribution of the Yi given the explanatory variables is multivariate
normal, i.e., that the ni independent residual or ‘error’ terms in ~i
have a multivariate normal distribution with zero mean vector, and
covariance matrix Xi. The latter is assumed to be a submatrix of
a T x T matrix, E, the elements of which are known functions of q
unknown covariance parameters, 8 , i.e., E = E(8). This model will
150 Design and Analysis of Clinical Trials

hold if the hypothetical complete-data, Y,’ can be written as:

where Yi,Xi and ~i are obtained by deleting rows from Yf,Xf and
€5, respectively.
Columns of each design matrix correspond to terms in the model
under consideration. The first column of each Xi,for example, is
a one-vector corresponding to the intercept term; other columns
may correspond t o dummy variables for grouping (between-subjects,
usually treatment) factors, and others to fixed and/or time varying
covariates.
Maximum likelihood estimates of p and 8 can be found by max-
imizing the log-likelihood L(8,p) given by:
. r n n

Numerical optimisation techniques are required to maximise L with


respect to p and 8. The details need not, however, concern us here
(they are given in Schluchter, 1988).
The likelihood approach outlined above is implemented in a num-
ber of software packages, e.g., BMDP5V and SAS PROC MIXED
(see Appendix), and such software enables the method to be applied
routinely to unbalanced longitudinal data from clinical trials.
Likelihood analyses of data that do not suffer from missing val-
ues are often (but by no means universally) robust to the use of
an incorrect model for the data. But with incomplete data such
analyses are likely to be more sensitive to model misspecification
since implicitly the data model is used to ‘fill-in’ for the missing val-
ues. So although valid likelihood inferences can be obtained when
the data contain non-informative missing values, the validity may
Multivariate Normal Regression Models for . . . 151

depend heavily upon using the correct model for the data. Con-
sequently, examining residuals and other diagnostics for detecting
model misspecification after fitting the proposed model becomes of
even greater importance than usual.
A standard maximum likelihood approach as described above is
known to produce biased estimators of the covariance parameters. In
many cases this problem is unlikely to be severe, but where it i s of
concern the method of restricted maximum likelihood or REML es-
timation, introduced originally by Patterson and Thompson (1971),
might be used. The details of REML estimation are given in Diggle
et al. (1994). Often the two estimation methods will give similar
results, but where they do differ substantially, those obtained from
REML are generally preferred.

6.5. NUMERICAL EXAMPLES

In this section, the model outlined above will be applied to a number


of examples, beginning with the trial of oestrogen patches for the
treatment of post-natal depression described in Chapter 5.

6.5.1. Oestrogen Patches in the Treatment of


Post-Natal Depression
Our first analysis of these data will involve the unrealistic assump-
tion that the repeated observations are independent. Under this
assumption, the model described in Section 6.3 is then simply that
of multiple regression. The graphical displays of these data given in
Chapter 5 suggest that there is a general decline in depression scores
over time and a difference in treatment level. Additionally, there is
some indication that the depression scores are beginning to ‘level-off’
by the end of the 12 months of the trial. This suggests considering
treatment group as a covariate, and including linear and perhaps
quadratic effects for time. In addition, the strong tracking observed
in the plots of standardised scores over time suggests that the pre-
randomisation measures should be incorporated into the model in
152 Design and Analysis of Clinical Dials

some way - we shall use the second of the pre-randomisation mea-


sures of depression as a further covariate. The model to be considered
can be written as follows:

where group is a dummy variable coding treatment received, 1 for


placebo and -1 for active, pre2 is the second pre-randomisation mea-
sure and time takes the values 1-6 depending on the visit involved.
The residual terms € i j k are assumed to normally distributed with
means zero and variance 8. Consequently the covariance matrix, z1,
of the repeated observations in each treatment group is assumed to
be:
z1 = eI (6.5)
where I is a 6 x 6 identity matrix.
The results from fitting this model are shown in Table 6.2. All
61 cases contribute values to the analysis despite only 45 of the
women having complete data. The effects of treatment group] the
pre-randomisation measure and linear time are all highly significant.

Table 6.2. Results from Likelihood Analysis of Oestrogen Patch


Data (assuming independence for the repeated observations).
Parameter Estimate Asymp.se z p-value
constant 7.087 2.057 3.446 0.0006
group 2.137 0.301 7.094 < 0.0001
linear time -2.090 0.823 -2.539 0.011
quadratic time 0.114 0.117 0.971 0.3315
baseline 0.477 0.079 6.020 < 0.0001
Approximate 95% confidence interval for treatment effect is 2 x 2.137 &
1.96 x 2 x 0.301 = (3.09,5.45)
The model has a single covariance parameter; the estimated value of this
parameter is 25.370 and its estimated standard error is 2.089.
The maximised log-likelihood for the model is -895.536, and Akaike’s in-
formation criterion takes the value -896.536.
Multivariate Normal Regression Models for . . . 153

The independence assumption made in the analysis reported


above is clearly far from the truth as Fig. 5.5 in the previous chapter
demonstrates, and so we now need to consider more realistic models
that allow for the repeated observations to be correlated. We shall
begin with a model that allows a particular structure, compound
symmetry, for the covariance matrix of the repeated measurements.
With this structure the covariance matrix in each treatment group
is assumed to be of the form:

!c=

i.e., a common covariance equal to 81 and a common variance equal


to 61 + 82. (As we shall see in Section 6.5, this structure arises from
assuming a simple mixed effects model for the data and is the basis of
the usual analysis of variance approach to the analysis of longitudinal
data.) The covariance matrix satisfying compound symmetry can be
written more concisely as follows:

where I is a 6 x 6 identity matrix and 1is a vector of ones.


The results from fitting a model with the same expected value
for the observations as given in Eq. (6.4), but with the compound
symmetry covariance structure, are shown in Table 6.3. Both the pa-
rameter estimates and their standard errors have changed from those
given in Table 6.2 although the conclusions remain largely the same.
The values of the log-likelihoods and Akaike’s information criterion
for the independence and compound symmetry models suggest that
the latter represents a great improvement.
154 Design and Analysis of Clinical Trials

Table 6.3. Results from Likelihood Analysis of Oestrogen Patch


Data (assuming compound symmetry for the covariance structure
of the repeated observations).

Parameter Estimate Asymp.se % p-value


constant 7.180 3.186 2.253 0.0242
group 1.995 0.543 3.673 0.0002
linear time -1.828 0.560 -3.263 0.0011
quadratic time 0.087 0.079 1.099 0.2719
baseline 0.460 0.145 3.174 0.0015
~

An approximate 95% confidence interval for the treatment effect is


(1.86,6.12).
The estimates and estimated standard errors of the model’s two covariance
parameters are 6 1 = 11.16 (1.029) and 6 2 = 14.399 (3.149).
The maximised log-likelihood is -831.764 and Akaike’s information crite-
rion is -833.764.

Allowing a compound symmetry structure for the correlations is


clearly more satisfactory than assuming that the repeated observa-
tions are independent of one another, but it remains an unrealis-
tic assumption for the oestrogen patch data in particular and for
longitudinal data in general. Both the constant variance and con-
stant covariance assumptions are unlikely in practice. Observations
taken close together in time are likely to have greater correlation
than those taken far apart. Figure 5.5 certainly suggests that this
is the case for the oestrogen patch data. One way to make certain
that the proposed model for the covariances is adequate is to fully pa-
rameterise x in the sense of considering N = T ( T +1)/2 parameters
81, 8 2 , . . . ,8 N corresponding to the unique variances and covariances
~ ~ 1 1 , 0 2 1. ., . , CTTT. Such a covariance matrix is usually referred to
as unstructured. Fitting such a model to the oestrogen patch data
gives the results shown in Table 6.4. Again the parameter estimates
and their standard errors change although the conclusions remain
largely the same. Comparing the values of Akaike’s criterion for this
model with the values corresponding to the two models considered
Multivariate Normal Regression Models for . . . 155

Table 6.4. Results from Likelihood Analysis of Oestrogen Patch


Data (allowing a fully parameterised covariance structure).
Parameter Estimate Asymp.se %
. p-value
constant 9.330 2.868 3.253 0.0011
group 2.047 0.486 4.211 < 0.0001
linear time -1.981 0.533 -3.718 0.0002
quadratic time 0.109 0.064 1.707 0.0879
baseline 0.364 0.129 2.820 0.0048
An approximate 95% confidence interval for the treatment effect is
(2.19,6.00).
The maximised log-likelihood is -781.343 and Akaike's information crite-
rion is -802.343.

previously shows that the unstructured covariance structure is to be


preferred.
Since a fully parameterised X will aZways provide a perfect de-
scription of the covariance matrix of the repeated observations, the
question arises as to why not use it on all occasions'? For the answer
we can return to the comments of Diggle (1988) already referred to in
the introduction to this chapter, namely that overparameterisation
can lead to inefficient estimation and potentially poor assessment of
standard errors of the parameters that characterise the treatment
mean profiles, i.e., those parameters that are of most interest in
the majority of applications. This is a problem which is likely to
be most critical when the number of observations is small relative to
the number of times at which observations are made.
When using the likelihood approach described above, the missing
values for subject i are imputed as follows:

yiz = XlZP + %2T;(ya - XZP) (6.8)


where y5 and Xg are partioned as follows:

y,'= ('a)
Yaz
andXf= (zz)
156 Design and Analysis of Clinical Trzals

with representing the missing values for this subject. The terms
yi2
212, k21 and 2 1 1 are appropriate submatrices of X(6). The esti-
mated missing values depend on the covariance structure assumed
for the repeated observations; this is illustrated in Table 6.5 which
shows the imputed missing values for the oestrogen patch data under
each of the covariance structures considered above.

Table 6.5. Imputed Missing Values for the Oestrogen Patch Data
(under three different models for the covariance structure of the
repeated observations).
Subject No. of Group Model
Missing Values Indep. CS Unstruc.
3 4 placebo 11.6 12.9 12.5
10.3 11.7 11.2
9.2 10.6 10.1
8.4 9.8 9.2
8 4 placebo 17.3 22.4 22.9
16.0 21.2 20.7
15.0 20.1 19.5
14.1 19.2 17.3
11 5 placebo 16.9 14.4 14.7
15.4 13.0 13.7
14.1 11.8 12.8
13.1 10.7 12.0
12.2 9.8 11.8
13 4 placebo 16.4 15.6 18.8
15.1 14.4 17.2
14.0 13.4 16.8
13.2 12.5 16.0
14 4 placebo 14.0 9.6 11.9
12.7 8.4 11.1
11.6 7.3 10.9
10.8 6.4 11.2
Multivariate Normal Regression Models for . . . 157

Table 6.5 (Continued)


Subject No. of Group Model
Missing Values Indep. CS Unstruc.
15 5 placebo 15.5 16.0 15.9
14.0 14.6 14.3
12.7 13.4 13.1
11.6 12.4 12.0
10.8 11.5 11.1
16 5 placebo 16.0 16.2 16.1
14.4 14.8 14.6
13.2 13.6 13.3
12.1 12.6 12.3
11.3 11.7 11.5
23 5 placebo 13.6 13.5 13.7
12.1 12.1 12.4
10.8 10.9 11.2
9.7 9.9 10.2
8.9 9.0 9.5
25 4 placebo 13.0 14.7 16.2
11.7 13.5 14.5
10.7 12.5 13.7
9.8 11.6 12.5
27 5 placebo 14.1 16.0 15.7
12.6 14.6 14.0
11.3 13.4 12.6
10.2 12.3 11.4
9.4 11.5 10.2
42 4 active 7.37 16.1 19.5
6.0 14.9 16.5
5.0 13.8 15.6
4.1 13.0 12.7
45 4 active 10.7 12.0 11.4
9.4 10.7 10.0
158 Design and Analysis of Clinical Dials

Table 6.5 (Continued)


Subject No. of Group Model
Missing Values Indep. CS Unstruc.
8.3 9.7 8.9
7.5 8.8 7.8
56 5 active 11.7 11.1 11.1
10.2 9.7 9.8
8.9 8.5 8.7
7.8 7.4 7.8
7.0 6.6 7.2
59 5 active 9.3 11.8 11.4
7.8 10.4 9.7
6.5 9.2 8.2
5.4 8.1 7.0
4.6 7.3 5.8
60 active 8.9 8.8 11.5
7.8 7.8 12.4
7.0 6.9 11.4
61 active 13.6 18.7 17.2
12.1 17.3 14.8
10.8 16.1 13.1
9.7 15.0 11.5
8.9 14.1 9.6

It was mentioned earlier that it is particularly important when


fitting models to data containing missing values to check model as-
sumptions. As with the usual version of multiple regression the most
useful diagnostics are the residuals, calculated as the differences be-
tween observed values and those predicted by the model. Here each
subject, i, will have a vector of residuals, ri, defined as follows:
A

ri = yi - Xip, (6.10)
Multivariate Normal Regression Models for ... 159

the individual elements of ri corresponding to the subject’s residual


at each time point. One possible method for displaying these resid-
uals is to first calculate each subjects Mahalanobis distances defined
thus:
- X,fi)
0; = (yi - xzfi)’2-1(yz (6.11)

If the assumptions of the model, i.e., normality and the form of


covariance matrix, are correct these distances have, approximately,
a chi-squared distribution with T degrees of freedom (remembering
that any missing values will have been imputed before the residuals
are calculated). Consequently, a chi-square probability plot of the
ordered distances should result in a straight line through the origin.
Departures from the assumptions will be indicated by depatures from
linearity in this plot.
The chi-squared plots for the Mahalanobis distances from the
three models fitted to the data from the oestrogen patch trial are
shown in Fig. 6.2. The plot corresponding to the independence model

Chbscuored plot of Mohalonobis distonces of


residuals from unstructured model
I 0 ”
0
0

oOOo
0
0

I I I

5 10 15

Chi-squore quonliler

Fig. 6.2(a)
160 Design and Analysis of Clinical Trials

Chi- squared plot of Mohalanobis distances of


residuais from compound symetry model
0

8ooo

I
I I

5 tO 15

Chi-square quonliies

(b)
Chi-squored plot of Moholanobis distances of
residuals from unstructured model
I 0
o o
0

ocloo
0

I I I

3 10 15

Chi-square quanliles

Fig. 6.2. Chi-squared probability plots of Mahalanobis distances from fit-


ting models to the oestrogen patch data. (a) Independence, (b) compound
symmetry (c) unstructured.
Multivariate Normal Regression Models for . . . 161

clearly departs from the expected y = z regression line but both the
compound symmetry and unstructured models lead to acceptable
plots.

6.5.2. Post-Surgical Recovery in Young


Children
In a comparison of the effects of varying doses of an anesthetic on
post-surgical recovery, 60 young children undergoing surgery were
randomised to receive one of four doses (15, 20, 25 and 30 mg/kg);
15 children were assigned to each dose. Recovery scores were assigned
upon admission to the recovery room and at 5, 15 and 30 minutes
following admission. The response at each of the four time points
was recorded on a six-point scale ranging from 0 (least favourable) to
6 (most favourable). In addition to the doseage, potential covariates
were age of patient (in months) and duration of surgery (in min-
utes). The data are shown in Table 6.6. In this case there are no
missing values. Plots of the individual response profiles by treatment
group and boxplots of the observations at each time point for each
treatment group are shown in Figs. 6.3 and 6.4.
These data will be used to illustrate what is known as a random
coeficients model in which both the intercept and slope (and possi-
ble higher order effects) of each individuals response profile can be
modelled as a random effect. Often, data which require a random
intercept term can be distinguished from those requiring fixed in-
tercepts by examining plots of the individual response profile (here
Fig. 6.3). If individual cases appear a s parallel lines, this suggests
a random intercept. In a data set requiring a fixed intercept but
random slopes, the marginal variance at the end time point will be
larger than at earlier time points - the lines spread out with in-
creasing time. A model with both random slope and intercept might
begin separated and would spread further, provided the association
between slope and intercept was positive. Although Fig. 6.3 is not
entirely convincing, here we shall fit a model with both random in-
tercepts and slopes.
162 Design and Analysis of Clinical Trials

Table 6.6. Anaesthesia Recovery Data.


Time (min after surgery)
Dose Patient Age Dur. 0 5 15 30
(mg/kg) (months) (min)
15 1 36 128 3 5 6 6
15 2 35 70 3 4 6 6
15 3 54 138 1 1 1 4
15 4 47 67 1 3 3 5
15 5 42 55 5 6 6 6
15 6 35 94 3 3 6 6
15 7 30 44 6 6 6 6
15 8 57 54 1 1 1 6
15 9 30 74 1 1 4 6
15 10 41 65 2 2 2 2
15 11 34 50 1 3 3 5
15 12 62 35 3 3 5 6
15 13 24 55 1 1 1 4
15 14 39 165 1 3 5 5
15 15 66 158 0 2 2 3
20 16 22 75 1 1 1 6
20 17 49 42 1 1 1 6
20 18 36 58 2 3 3 6
20 19 43 60 1 1 2 3
20 20 23 64 5 6 6 6
20 21 30 46 1 1 2 4
20 22 9 114 6 6 6 6
20 23 14 50 4 4 6 6
20 24 2 95 1 4 5 5
20 25 50 125 1 2 2 5
20 26 26 127 6 6 6 6
20 27 40 173 0 0 0 4
20 28 12 110 3 6 6 6
20 29 42 47 1 1 5 6
20 30 18 97 2 2 3 5
Multivariate Normal Regression Models for . .. 163

Table 6.6 (Continued)


Time (min after surgery)
Dose Patient Age Dur. 0 5 15 30
(mg/kg) (months) (min)
25 31 26 103 1 1 0 3
25 32 28 89 3 6 6 6
25 33 41 51 2 3 4 4
25 34 46 93 1 1 5 6
25 35 37 45 2 3 6 6
25 36 28 68 6 6 6 6
25 37 37 35 3 5 6 6
25 38 60 54 2 3 3 6
25 39 60 55 1 1 1 3
25 40 38 78 0 2 6 6
25 41 47 118 0 0 0 0
25 42 38 98 1 1 1 4
25 43 23 58 1 2 6 6
25 44 56 190 1 1 1 1
25 45 31 125 0 3 5 6
30 46 46 72 4 6 6 6
30 47 38 85 2 4 6 6
30 48 59 54 4 5 5 6
30 49 16 100 1 1 1 1
30 50 65 113 2 3 3 5
30 51 53 72 3 4 4 6
30 52 50 70 0 5 5 5
30 53 13 85 0 0 0 4
30 54 17 25 0 0 0 0
30 55 70 53 1 1 1 4
30 56 13 45 0 0 4 6
30 57 60 41 1 1 4 6
30 58 12 61 1 1 4 6
30 59 27 61 3 5 5 6
30 60 56 106 0 1 1 3
164 Design and Analysis of Clinical Dials

Dose 15rng Dose 20rng

I 1 1 I , 1

0 5 15 30 0 5 15 30
Tme(mins) Time(rnins)

Dose 25mg Dose 30rng

w -

b -

N N -

0 0 -
I , 1 I , 1

0 5 15 30 0 5 15 30

Time(mins) Time(rnins)

Fig. 6.3. Individual response profiles for post-surgical recovery data.


Multivariate Normal Regression Models for 165

Dose 15mg Dose 20mg

w - T T

In-

* -

m -

N - w

- - LL.I

0 A 0
I

Dose 25rng Dose 30mg

m -

N -

7 -

Fig. 6.4. Box-plots of surgical-recovery data.


166 Design and Analysis of Clinical Traals

We assume that the recovery scores for the kth subject at the j t h
time point and the ith drug dose, y i j k , are given by:

where a i k and bik are random intercept and slope for subject k having
dose i, and ~ i j are
k random error terms. The t j are the times at which
the recovery scores are measured. The random intercept and slope
terms are assumed to have a multivariate normal distribution:

Likewise, the random error terms are assumed to have the following
distribution:

The random error terms are assumed to be independent of the ran-


dom intercepts and slopes.
The model can be written more concisely as:

(6.15)

z = ( ;1 ’.).
30 (6.16)

Under this model, the expected values of the recovery scores are:

E ( Y i j k ) = Qi + pztj (6.17)
and the covariance matrix of y i k has the following random effects
structure:
z: = ZQZ‘ + a21 (6.18)
Multivariate Normal Regression Models for 167

Table 6.7. Results of Likelihood-based Anal-


ysis of Recovery Data (assuming random
coefficients model).

Term df chi-square p-value


treatment 3 1.27 0.736
linear time 1 165.88 < 0.001
treatment x time 3 0.10 0.992

The parameter estimates obtained from using BMDP 5V are shown


in Table 6.7. To assess differences between the drug groups in in-
tercepts and slopes, W d d ’ s test can be used; this assesses whether
a particular subset of the parameters, Pm, is zero. The form of this
test is:
w =@p1jm (6.19)

where V is the covariance matrix of the estimates in the subset of


interest. If all m parameters in the subset are zero, then W has a
chi-squared distribution with m degrees of freedom.
The results of this test for intercepts and slopes are also shown
in Table 6.7. There is no evidence of any difference between the
different drug dose groups.

6.6. ANOVA A N D MANOVA


APPROACHES TO THE ANALYSIS OF
LONGITUDINAL DATA

One of the most common types of analysis for longitudinal data, par-
ticularly in psychiatry and psychology, remains either univariate or
multivariate analysis of variance. The ANOVA approach is usually
formulated in terms of a mixed-effects model which includes a ran-
dom subject effect, this allowing the repeated observations to have
the compound symmetry covariance structure. Such a model for the
168 Design and Analysis of Clinical Trials

oestrogen patch data takes the form:

where yijk denotes the depression score of subject k in treatment


group i on the j t h visit, p represents an overall mean effect, cq, i =
1 , 2 represent the effect of treatment, pj, j = 1,.. . , 6 represent time
effects, and yij denotes the groupxtime interaction. The term Tk is
used for the subject effect specific to subject k, and € i j k represents the
usual residual or ‘error’ terms. The terms a , P and y are considered
fixed effects and both 7 and E are assumed to be random effects
having normal distributions with zero means and variances 81 and
02. Such a model implies that the covariance matrix of the repeated
observations is as shown in Eq. (6.6).
The MANOVA route removes the constraints on the covariances
implied by compound symmetry and allows the covariance matrix
of the repeated observations to be fully parameterised as described
in the previous section. The problems with assuming compound
symmetry and in some situations with allowing an unstructured co-
variance matrix have been described in Section 6.4. These problems
make the use of either ANOVA or MANOVA less than ideal in most
situations, although for longitudinal data containing no missing val-
ues the latter is unlikely to be seriously misleading. But when the
data are affected by dropouts, the application of either ANOVA or
MANOVA to complete cases only (still the most usual approach to
missing values) should be avoided. If the missing values are MCAR,
both approaches do not use the available data as efficiently as they
should; much more serious, however, is that if the data are MAR,
analysing complete cases only might lead to misleading inferences.

6.7. INFORMATIVE DROP OUTS

The analyses carried out on the oestrogen patch data in Section 6.4
are only strictly valid if the missing observations caused by the
Multivariate Normal Regression Models f o r . . . 169

women who dropped out are non-informative. Two questions there-


fore arise:

0 How do we tell whether or not the missing observations are


ignorable?
0 How do we analyse the data if the missing values are informa-
tive?

A number of authors have considered these questions including,


Laird (1988), Diggle and Kenward (1994) and Diggle (1998). (Rid-
out, 1991, considers the slightly simpler problem of testing for com-
pletely random dropouts and proposes a method based on logistic
regression.) Diggle (1998) proposes an informal approach to assess-
ing whether missing values are likely to be informative or otherwise,
involving consideration of the mean profiles of cohorts of completers
and those who drop out at particular visits, ignoring the different
treatment groups. Figure 6.4 illustrates this proposal on the oestro-
gen patch example; the completers cohort has steadily decreasing
depression scores whereas the cohorts of dropouts have increasing
values. (The number of missing values in this example is small so
that some of the points plotted in Fig. 6.4 are based on just a few
subjects; the plot does however serve to illustrate the general point
being made.)
A more formal approach is that suggested by Diggle and Ken-
ward (1994) who propose a modelling framework for longitudinal
data with informative dropouts, in which random or completely ran-
dom dropouts are included as explicit models. A brief technical ac-
count of the model is given in Table 6.8, but the essential feature is
a logistic model for the probability of dropping out, in which the ex-
planatory variables can include previous values of the response vari-
able as in the model described above, but in addition can include the
unobserved value at dropout as a latent variable. In other words, the
dropout probability is allowed to depend on both the observed mea-
surement history and the unobserved value at dropout. The model
is implemented in the software OSWALD (see Appendix).
170 Design and Analysis of Clinical Dials

Table 6.8. Diggle and Kenward Model for Dropouts.


Let Y * represent the complete vector of intended measurements and
t = [ t l , t 2 , .. . , t,] the corresponding set of times at which measurements
are taken.
Let Y = [Yl, Y2,. . . , Y,] denote the vector of measurements, with missing
values coded as zero.
Y* and Y coincide as long as the individual remains in the study, im-
plying

Yk = Y { , k = 1 , 2 ,...,D - 1
= 0 k>D

where D is a random variable such that 2 5 D 5 n identifies the dropout


+
time and D = n 1 identifies no dropout.
The distribution of Y *is assumed to be multivariate normal with prob-
ability density function f *(y;p, e),where p and e, respectively param-
eterise the mean and covariance structure of Y * .
We assume that the dropout process is such that the probability of
dropout at time t d depends on the history of measurement up t o and
including td.
For each k , let HI, = [yl,. . . , yk-11 represent the observed history up t o
time tk-1. The model for the dropout process is then:

P r ( D = dlhistory) = pd(Hd, y;; 4)


So the dropout probability depends on both the observed measurement
history H,j, and the unobserved value yC;. In addition, the probability
depends on a set of parameters 4.
The equations that determine the joint distribution of Y , and hence the
likelihood function of the parameters, p, 8 and 4 are:

Pr(Yk = OIHI,,Yk-1 = 0 ) = 1
Multivariate Normal Regression Models for . . . 171

Table 6.8. (Continued)

where fl(y1H1,;p, 8) denotes the conditional alternate Gaussian pdf of


y i given HI, and f k ( y ( H ~p
, ;, 8 , 4 ) the conditional pdf of YI, given Hk.
0 A linear logistic model is used for p~,(Hk, y; 4).
c From there the likelihood function for p, 8 and 4 can be constructed.
(This account is an abbreviated version of that given in Diggle and Ken-
ward, 1994, where readers are referred to for more detail.)

Table 6.9. Results from DiggleKenward Model for Dropouts


Oestrogen Patch Data (assuming random drop-out based on
1 previous observation).
Analysis method: Maximum likelihood (ML)
Maximised likelihood:
[l]- 1295.658
Mean Parameters:

(Intercept) Group Time Pre2


PARAMETER -1.218 -1.751 -0.572 0.807
STD. ERROR 20.58 0.356 0.071 0.094

Dropout parameters:

(Intercept) Y.d y.d - 1


-3.941 0 0.103

To illustrate the Diggle-Kenward model for dropouts, the method


will be applied to the oestrogen patch example. Part of the outputs
from OSWALD corresponding to models for, MAR and informative
dropouts are shown in Tables 6.9 and 6.10. The likelihood ratio
test for the additional parameter in the informative dropout model
indicates that the parameter is not zero. This finding might appear to
throw into doubt the previously reported likelihood analyses of these
data in which the dropouts were assumed to be non-informative.
172 Design and Analysis of Clinical Trials

Table 6.10. Results from Diggle-Kenward Model for Dropouts


Oestrogen Patch Data (assuming informative d r o p o u t ) .

Maximised likelihood:
[l]- 1291.848
Mean Parameters:

(Intercept) Group Time Pre2


PARAMETER 10.198 -2.335 -0.670 0.309
STD. ERROR NA NA NA NA
Dropout parameters:

(Intercept) Y.d y.d - 1


~~ ~

-4.797 0.022 0.131


0 LR test of additional parameter: 7.62 1d.f.
(The standard errors of the mean parameters for the informative
dropout model are not available from OSWALD.)

However, the parameters characterising the mean profiles tell largely


the same story in both models. Certainly the likelihood analysis as-
suming only MAR is less likely to suffer from bias than the alternative
of using only complete cases and implicitly assuming MCAR.
The Diggle-Kenward model represents a welcome addition to the
methodology available for analysing longitudinal data in which there
are dropouts. But as with any new modelling framework, ques-
tions need to be asked about its adequacy in practical situations.
Matthews (1994), for example, makes the point that if there are many
dropouts, the proposed model can be applied, but questions whether
many statisticians would feel happy to rely on technical virtuosity
when 60% of the data are absent. Alternatively, if the proportion of
dropouts is low, then much less can be learnt about the dropout pro-
cess, leading to low power to discriminate between dropout processes.
Skinner (1994) suggests that the longitudinal data remaining after
dropout may, by themselves, contain very little information about the
Multivariate Normal Regression Models for . . . 173

‘informativeness’ of the dropout and concludes that external infor-


mation about the dropout mechanism be sought and used. A further
possible problem identified by Troxel et al. (1998) is that bias that
results from assuming the wrong type of missingness mechanism may
well be more severe than the bias that results from misspecification
of a standard full maximum likelihood model.
Despite these and other reservations made in the discussion of
Diggle and Kenward’s paper, their proposed model does open up
the possibility of some almost routine, detailed investigation of the
dropout process.

6.8. OTHER METHODS FOR THE


ANALYSIS OF LONGITUDINAL DATA
There are a number of methods of analysis that allow for more gen-
eral covariance matrices to be specified. One, very little used within
the clinical trials field, is structural equation modelling (SEM). As
typically presented, SEM typically start by a detailed consideration
of the appropriate structure for the covariance matrix and only later
proceeding to consider the structure of the means that is typically
the main focus for the trial analyst. However, a common limitation of
SEM software is their limited ability t o deal with missing data, typi-
cally approached through the fitting of the model to multiple groups,
one corresponding to each group of subjects with a common pattern
of missing data (see Dunn, Everitt and Pickles, 1993). Rather more
practicable are programs such as Mx (Neale, 1995) that perform what
is described as full-information maximum-likelihood estimation, that
fit the model at the individual subject level if required.
A second increasingly important method uses software for mul-
tilevel modelling (e.g., MLWiN - see Appendix). These, too, are
capable of fitting models with error covariance matrices with struc-
tures like those already described. However, in addition, the appro-
priate use of dummy variables with random coefficients allows for
heteroscedasticity in variances and covariances between treatment
groups.
174 Design and Analysis of Clinical lhals

A novel approach to the analysis of longitudinal data is sug-


gested by Tango (1998). Tango postulates that each treatment group
consists of a mixture of several distinct latent profiles, a situation
he models using a finite mixture model with normal components
(see Everitt, 1996). The effect of treatment is characterised by the
slope of the latent profiles and the mixing proportions of these la-
tent profiles. An EM algorithm is used for parameter estimation.

Means for dropouts and completers

r I I I I I
v1 v2 v3 v4 v5 V6

Visit

Fig. 6.5. Means for completers and cohorts of subjects dropping out a
particular times for oestrogen patch data.
Multivariate Normal Regression Models for . . . 175

Tango illustrates his proposal on data from a randomised double


blind trial of two treatments for chronic hepatitis, a new treatment
A, and the standard treatment B. The response variable used
was log-transformed serum levels of glutamate pyruvate transami-
nase (GPT), measured at baseline and at one week intervals up to
four weeks. A five-group mixture model was deemed optimal, the
groups being labelled, ‘greatly improved’, ‘improved’, ‘unchanged’,
‘worsened’, ‘greatly worsened’. A patient’s group membership was
determined by the maximum value of their estimated posterior prob-
abilities of belonging to each group. The profiles of all the 124 pa-
tients and of the patients in each of the derived groups are shown in
Fig. 6.5. The composition of the two-treatment groups in terms of
each type of patient was as follows:

Treat- Greatly Improved Unchanged Worsened Greatly Total


ment Improved Worsened

A 3 20 17 14 8 62
B 2 13 34 11 2 62

Compared with the standard treatment B, the treatment A has:

0 the same proportion of ‘greatly improved’,


0 higher proportion of ‘improved’,
0 higher proportion of ‘worsened’,
0 higher proportion of ‘greatly worsened’,
0 lower proportion of ‘unchanged’.
Tango suggests that these characterisations of the efficacy of
treatments might be medically important especially for finding the
key baseline factors to discriminate responders from non-responders.

6.9. SUMMARY
For longitudinal data where the response variable has a normal distri-
bution, the regression model described in Section 6.4 can be applied
176 D e s i g n a n d A n a l y s i s of Clinical Trials

1 improved

Fig. 6.6. Individual profiles for all patients and for mixture modelling-
derived 'types'.

routinely] using software such as PROC MIXED in SAS and


BMDP5V. When the data are unbalanced but the missing values
can be assumed to be MCAR or MAR, the estimation of the param-
eters in such models by maximum likelihood provides a completely
satisfactory method of analysis and one that is much to be preferred
over the commonly used application of univariate or multivariate
analysis of variance to complete cases.
Although the missing-at-random assumption made in the like-
lihood approach described in this chapter is often sensible, Little
and Yau (1996) give some examples where it is highly questionable.
In such cases, the Diggle-Kenward model for informative dropouts
can be applied using the OSWALD package, although at present
this should perhaps be regarded as primarily experimental. In addi-
tion, Little (1995) suggests that the model may be highly sensitive
Multivariate Normal Regression Models for . . . 177

to model misspecification and in Little and Lau (1996) describes an


alternative approach to the dropout problem. Now values of the out-
come after dropout are imputed using models for the missing data
that condition on all relevant observed data, and in particular on
information about treatments actually administered (as opposed to
randomised treatment). Following imputation an intention-to-treat
analysis based on the treatment as randomised is applied to the im-
puted data.
Finally, it must be remembered that the proportion of dropouts
may, for many trials, be regarded as a measure of the quality of
the data produced by the trial. And Lavori’s seemingly self-evident
statement, “it is always better to have no dropouts” (Lavori, 1992);
still has considerable appeal.
CHAPTER7

Models for Non-Normal


Longitudinal Data
from Clinical Trials

7.1. INTRODUCTION

In Chapter 4 we gave a brief description of generalised linear mod-


els and showed how they have unified regression analysis for discrete
and continuous independent response variables. An obvious question
is how might such models be adapted t o be suitable for correlated
response variables, in particular the repeated measurements taken
in a longitudinal study? The multivariate regression model for cor-
related normally distributed response variables found in Chapter 6
provides the first part of the answer, and in this chapter the extension
to non-normal responses, in particular, categorical responses, will
be considered.
In the linear model for Gaussian data in Chapter 6 estimation
of the regression parameters needed to take account of the correla-
tions in the data, but their interpretation was essentially independent
of the correlation structure. One potentially troublesome aspect of
modelling non-normal longitudinal data is that this independence
of estimation and interpretation no longer always holds. Different
assumptions about the source of correlations between the observa-
tions can lead to regression coefficients with distinct interpretations

178
Models for Non-Normal Longitudinal Data from Clinical Dials 179

as we shall see later. The three principal approaches to introduc-


ing correlations are marginal, random eflects and transition models.
Each type will now be considered relatively briefly; fuller accounts
are available in Diggle, et al. (1994).

7.2. MARGINAL MODELS

Longitudinal data can be considered as a series of cross-sections, and


marginal models for such data use the generalised linear model dis-
cussed in Chapter 4,to fit to each cross-section. In this approach, the
relationship of the marginal mean and the explanatory variables is
modelled separately from the within-subject correlation. Specifically,
a marginal model makes the following assumptions:

and CY a set of parameters.

(The nomenclature is that introduced for GLMs in Chapter 4


and for longitudinal data in Chapter 6, with pij now representing
the expected value of the observation on subject i at time t j , i.e.,
Xj.)
The first two of these assumptions are exactly the same as those
made in Chapter 4 for generalised linear models, suitable when only
a single response is observed on each subject. The marginal regres-
sion coefficients have the same interpretation as coefficients from a
cross-sectional analysis, and marginal models are natural analogues
for correlated data of generalised linear models for independent data.
Before considering how the parameters in such models can be es-
timated, it might be helpful to look at a simple example. For this we
shall use the respiratory status data introduced in Chapter 5. These
data consist of binary observations of whether a patient had good or
poor respiratory status at baseline and at each of four post-treatment
180 Design and Analysis of Clinical Dials

visits. Two treatment groups were involved and in addition two fur-
ther covariates, sex and age, were available together with a binary
indicator for centre. One possible logistic marginal model is given
by:

0 Corr(xj,yZk) = 0

where treatmenti is a dummy variable indicating the treatment group


of subject i.
Here the parameter exp(P1) is the odds ratio of good respira-
tory status in the two treatment groups. Since exp(P1) is a ratio
of population frequencies, it is referred t o as a population-averaged
parameter. If all the individuals in the same treatment group have
the same probability of good respiratory status, the population fre-
quency is the same as the individual’s probability. When, however,
there is heterogeneity in the probability in a treatment group, the
population frequency is the average of the individual values. (This
point will be discussed further later in this chapter.)
By setting the correlation between pairs of observations to zero,
we are simply ignoring the likely lack of independence of observations
over cross-sections. This is an example of what Davis (1991) refers
to as the independence working model approach. Although such an
approach can yield consistent estimates of the regression parameters
in many circumstances, the standard estimators of the variances and
covariances of these estimates will not be consistent. This problem
can be overcome in a variety of ways, but perhaps the simplest is to
make use of an extension of the robust covariance matrix estimator
of Huber (1967) and White (1982) (see Chapter 4) that makes it
suitable for correlated data.
Fitting the independence working model does not need any spe-
cialised software since standard GLM or logistic regression packages
can be used. Response measures are organised one per record with
each subject contributing as many records as repeated measures.
Models for Non-Normal Longitudinal Data from Clinical Trials 181

Table 7.1. Parameter Estimates from an Independence


Working Model: Logistic Regression Fitted to the Respi-
ratory Status Data.
Covariate Estm. Regres. Standard Errors
Coefficient Classical Robust Bootstrap
Treatment 1.267 0.235 0.349 0.383
Time -0.078 0.099 0.082 0.083
Sex 0.137 0.294 0.443 0.482
Age1100 -0.189 0.883 1.305 1.354
Baseline 1.849 0.240 0.349 0.395
Centre 0.651 0.238 0.355 0.372
The bootstrap SE is based on 500 replicates.

The operational simplicity of this approach is hard to exaggerate


and Table 7.1 gives the parameter estimates from fitting the follow-
ing logistic regression model to the respiratory status data given in
Table 5.9.

logit& = 1) = Po + Pi treatmenti + ,& timeij + ,& sexi


+ P4 agei + /&, centrei + ,& baselinei . (7.1)
Three forms of standard error are shown together with their associ-
ated test statistic; the classical standard error that is based on the
unlikely assumption that the four repeated observations on a sub-
ject are independent, the ‘robust/sandwich’ standard error, and a
bootstrap standard error. The bootstrap resampling recognised the
repeated measures structure by resampling subjects, each with a set
of responses, rather than using single responses. The bootstrap esti-
mate is far closer to the robust/sandwich estimate than the classical
estimate, which, for between subjects effects, are much too small.
But in the case of the within subjects effects for the time trend over
assessment visits, it is the robust and bootstrap estimates that are
smaller. Assuming independence is not always anti-conservative.
182 Design and Analysis of Clinical Trials

It is important to keep in mind that what is being estimated by


the fitted model is the cross-sectional relationship between variables.
Table 7.2 shows the sample frequencies over the joint distribution
of the baseline and four trial responses (2 times 16 possible binary
response sequences), together with the frequencies predicted by the
independence working model given in Eq. (7.1). The other entries
in this table will be discussed later. Whilst the time-by-time marginal
totals are well fitted by this model, the joint distribution reflected
by these response sequences is very poorly fitted. Typical of such
data, the sequences with a preponderance of one type of response
are more common than expected under the assumption of indepen-
dence, and the sequences with alternating responses considerably
fewer. This should come as no surprise since the model was not
fitted to the joint distribution, only to the margins.

Table 7.2. Sample Frequencies and Predicted Frequencies for


Respiratory Status Data.

Response Observed Independence Transition Random Effects


Frequency Working Model Model Model

Baseline 0
0000 25 13.3 14.3 21.9
0001 2 4.3 4.1 3.1
0010 4 4.6 4.6 3.5
001 1 0 2.3 3.0 1.4
0100 2 5.0 5.6 4.0
0101 0 2.5 2.9 1.6
0110 2 2.7 3.7 1.8
0111 4 2.1 4.1 1.7
1000 3 5.4 4.2 4.5
1001 0 2.7 1.7 1.8
1010 1 2.9 2.0 2.0
1011 2 2.2 1.8 1.9
1100 2 3.2 2.2 2.2
Models for Non-Normal Longitudinal Data from Clinical Trials 183

Table 7.2. (Continued)


Response Observed Independence Transition Random Effects
Frequency Working Model Model Model

1101 3 2.4 1.6 2.2


1110 1 2.6 2.1 2.5
1111 10 2.7 3.2 5.0
Baseline 1
0000 1 0.3 0.2 0.4
0001 0 0.5 0.3 0.4
0010 0 0.6 0.3 0.5
0011 3 1.2 0.6 0.7
0100 1 0.6 0.6 0.6
0101 0 1.3 0.9 0.8
0110 1 1.4 0.9 0.9
0111 1 4.0 2.4 2.6
1000 4 0.7 1.1 0.6
1001 2 1.4 1.8 0.9
1010 1 1.5 1.6 1.0
1011 3 4.3 3.9 3.0
1100 0 1.6 3.1 1.2
1101 1 4.6 5.7 3.4
1110 5 5.0 5.5 3.8
1111 27 21.0 21.2 29.2

7.2.1. Marginal Modelling using Generalised


Estimating Equations
The independence working model used above can be estimated using
the approach outlined for generalised linear models in Chapter 4. But
for more complex marginal models in which we wish to take advan-
tage of the correlation between pairs of observations to obtain more
efficient estimates of the regression parameters, we need to consider
a new procedure introduced by Liang and Zeger (1986), and known
as generalised estimating equations (GEE). This approach may be
184 Design and Analysis of Clinical Trials

viewed as a multivariate extension of the generalised linear model


and the quasi-likelihood method (see Chapter 4). The use of the
latter leads to consistent inferences about mean responses without
requiring specific assumptions to be made about second and higher
order moments. In this way intractable likelihood functions with pos-
sibly many nuisance parameters, in addition to p and a,are avoided.
A brief account of GEE is given in Table 7.3. More detailed accounts
are available in Liang and Zeger (1986) and Zeger, Liang and Pren-
tice (1988). (Software for fitting GEE models is discussed in the
Appendix.)

Table 7.3. Generalised E s t i m a t i n g Equations.


0 In the absence of a likelihood function, Liang and Zeger (1986) show
that the parameters of a generalised linear model for longitudinal data,
p and a,can be otained from a multivariate analogue of the quasilike-
lihood mentioned briefly in Chapter 4.
0 The generalised estimating equation is given by:

U ( P , a )= c(y ) ’ v ; l ( y i
n

i=l
- pi@))= 0

0 V; is assumed to be of the form:

where Ai is a diagonal matrix of variances, the elements of which de-


pend on the elements of p i , and Ri(a) is a working correlation matrix.
0 Some possibilities for R i ( a ) are:
(a) An identity matrix leading to the independence working model in
which the generalised estimating equation reduces to the univariate
estimating equation given in Chapter 4, obtained by assuming that
the elements of yi are independent.
(b) An exchangeable correlation matrix with a single parameter similar
to that described in Chapter 6. Here corr(Y,j, Y i k ) = a.
(c) An AR-1 autoregressive correlation matrix, also with a single pa-
rameter, but in which corr(Y,j, &) = alk-jl, j # Ic.
Models for Non-Normal Longitudinal Data from Clinical Rials 185

Table 7. ( Continued )
~ ~~ ~

(d) An unstructured correlation matrix with T ( T - 1)/2 parameters in


which corr(yZ, , x k ) = L Y , ~ .
0 In cases of possible overdispersion, the scale parameter 4 can again be
estimated by the moments estimator mentioned in Chapter 4, using
observation specific Pearson residuals, i,,.
One version of GEE also uses moment estimators for the correlation
parameters. For the exchangeable model, for example,

The estimation process then consists of an iteration between iterative


weighted least squares estimation of the regression parameters for a
given estimate of the correlation parameters, followed by a recalculation
of the latter based on the residuals from the current estimates of the
former.
Other approaches t o GEE estimation are described in Zhao and Prentice
(1990) and Prentice and Zhao (1991).
The GEE method produces consistent estimators for p and its
covariance matrix without requiring: (a) specification of the joint dis-
tribution of the repeated observations for each individual; (b) correct-
specification of the correlation structure in U(p, a).

Table 7.4 presents results from logistic regression models fitted to


the respiratory status data using GEE under a variety of assumptions
about the correlational structure. It is sufficient for our purposes t o
present results only for the treatment effect of main interest, and
the trend over occasions - the only variable that varies within sub-
jects. Both classical and robust standard errors are given, although
we should now expect the differences between these two standard
error estimators to be smaller since, unlike the independence work-
ing model used earlier, these models all make at least a plausible
assumption as to the correlation structure (marked differences might
have suggested possible model misspecification) .
186 Design and Analysis of Clinical Trials

Table 7.4. Results from Fitting Logistic Regression Model to


Respiratory Status Data under Various Assumptions about Cor-
relational Structure.
Classical Robust
Covariate Model Estimate SE ztest SE atest
Treatment Exch 1.256 0.331 3.79 0.348 3.61
AR-1 1.205 0.308 3.91 0.349 3.45
Unstr 1.239 0.330 3.76 0.347 3.58
Time Exch -0.078 0.081 -0.97 0.082 -0.95
AR-1 -0.097 0.101 -0.96 0.083 -1.18
Unstr -0.086 0.085 -1.01 0.082 -1.05

Table 7.5 gives the lower triangle of the estimated correlation


matrices from each of these models. The estimated correlations from
the unstructured model look to be closer to those of an exchangeable
structure than the autoregressive (AR-1) structure. In the latter
model, the correlation between pairs of observations are expected to
decline with separation in time much more quickly than they actually
appear to do, whereas the exchangeable model assumes that this
correlation remains constant. This greater similarity, in this case, of
the unstructured and exchangeable models as compared to the AR-1
model, is also reflected in the estimated regression coefficients and
standard errors.
Nonetheless all the models give similar findings, with there being
little doubt as to the presence of a treatment effect and there being
little trend over assesment visit.
Table 7.6 presents some well known data originally presented
by Thall and Vail (1990) involving a clinical trial of treatment for
epilepsy in which a total of 59 patients were randomly allocated to
either a placebo or the active treatment, namely the drug progabide.
In this trial, the response variable was a count of the number of
seizures within four successive intervals of two weeks. A baseline
Models for Non-Normal Longitudinal Data from Clinical Trials 187

Table 7.5. Estimated Correlation Matrices for the


Respiratory Status Data.
(a) Exchangeable

-
R= [ 1.oo
0.33(0.18) 1.00
0.33(0.18) 0.33(0.18) 1.00
0.33(0.18) 0.33(0.18) 0.33(0.18) 1.00

(b) AR-1

0.15 0.38 1.00


0.06 0.15 0.38 1.00

( c ) Unstructured

0.32 1.00
R=
0.20 0.42 1.00
\0.29 0.34 0.38 1.00)

measure of the response was also taken and, in addition, the age of
each patient was recorded. Figure 7.1 is a plot of the log of the total
number of seizures (plus one) during the trial against the log of the
baseline seizure count. Except for subject 58, with an unusually low
number of seizures during the trial, the relationship looks plausibly
linear on this log scale. The plot also suggests that subject 49 is
likely to have high influence.
The difference in mean log-total for placebo and treatment
groups (3.21 - s.e. 0.16 and 2.80 - s.e. 0.19) of 0.41 was not
significantly different from zero using a simple t-test. Of course,
controlling for baseline differences improves efficiency. However, A
188 Design and Analysis of Clinical Rials

Table 7.6. Data from a Clinical Trial of Patients Suffering


from Epilepsy.

ID
- Y1 Yz Y3 Y4 Treatment Baseline Age
1 5 3 3 3 0 11 31
2 3 5 3 3 0 11 30
3 2 4 0 5 0 6 25
4 4 4 1 4 0 8 36
5 7 18 9 21 0 66 22
6 5 2 8 7 0 27 29
7 6 4 0 2 0 12 31
8 40 20 23 12 0 52 42
9 5 6 6 5 0 23 37
10 14 13 6 0 0 10 28
11 26 12 6 22 0 52 36
12 12 6 8 4 0 33 24
13 4 4 6 2 0 18 23
14 7 9 12 14 0 42 36
15 16 24 10 9 0 87 26
16 11 0 0 5 0 50 26
17 0 0 3 3 0 18 28
18 37 29 28 29 0 111 31
19 3 5 2 5 0 18 32
20 3 0 6 7 0 20 21
21 3 4 3 4 0 12 29
22 3 4 3 4 0 9 21
23 2 3 3 5 0 17 32
24 8 12 2 8 0 28 25
25 18 24 76 25 0 55 30
26 2 1 2 1 0 9 40
27 3 1 4 2 0 10 19
28 13 15 13 12 0 47 22
29 11 14 9 8 1 76 18
30 8 7 9 4 1 38 32
Models for Non-Normal Longitudinal Data from Clinical Trials 189

Table 7.6. Data from a Clinical Trial of Patients Suffering


from Epilepsy.

ID y1 yz y3 y4 Treatment Baseline Age


31 0 4 3 0 1 19 20
32 3 6 1 3 1 10 30
33 2 6 7 4 1 19 18
34 4 3 1 3 1 24 24
35 22 17 19 16 1 31 30
36 5 4 7 4 1 14 35
37 2 4 0 4 1 11 27
38 3 7 7 7 1 67 20
39 4 18 2 5 1 41 22
40 2 1 1 0 1 7 28
41 0 2 4 0 1 22 23
42 5 4 0 3 1 13 40
43 11 14 25 15 1 46 33
44 10 5 3 8 1 36 21
45 19 7 6 7 1 38 35
46 1 1 2 3 1 7 25
47 6 10 8 8 1 36 26
48 2 1 0 0 1 11 25
49 102 65 72 63 1 151 22
50 4 3 2 4 1 22 32
51 8 6 5 7 1 41 25
52 1 3 1 5 1 32 35
53 18 11 28 13 1 56 21
54 6 3 4 0 1 24 41
55 3 5 4 3 1 16 32
56 1 23 19 8 1 22 26
57 2 3 0 1 1 25 21
58 0 0 0 0 1 13 36
59 1 4 3 2 1 12 37
190 Design and Analysis of Clinical Dials

0
0
0
0
0 00
0 0 o o
00 0
0
O
0 0
O38
0 0

o"o"go 02000 0

0 00 O O

0 0
0 0 0
0
0

0
1 I I 1 1 I I
2.0 2.5 3.0 3.5 4.0 4.5 5.0

Log(base1ine seizure count+l)

Fig. 7.1. Plot of log of total number of seizures (plus one) against log of
the baseline seizure count for epilepsy data in Table 7.6.

Cook-Heisberg test identified heteroscedasticity in a regression of


log-total on treatment group, log-baseline and age, suggesting the
need to use the Huberized/robust/sandwich (heteroscedasticity con-
sistent) parameter covariance estimator in making inference about
effects in such a model. This suggested a significant treatment effect
conditional upon covariates of -0.38 with a 95% confidence interval
of (-0.71, .-0.05).
We have already noted in Chapter 4 how regression analyses
of log transformed data can be distorted by obervations with re-
sponses at or close to 0. Subject 58 is clearly one such case. Per-
haps the natural model to analyse count data such as these is some
form of GLM based on a Poisson error distribution. However, fit-
ting a GLM with covariates for treatment, age and log-baseline,
Models f o r Non-Normal Longitudinal Data from Clinical Trials 191

Table 7.7. GEE Poisson Regression Estimates for the Epilepsy


Data.
~~ ~

Covariate Model Estimate SE %test


Treatment (1) Indep with robust -0.031 0.193 -0.163
(2) Exch -0.025 0.071 -0.361
(3) Exch with overdisp -0.025 0.155 -0.164
(4) Exch with overdisp
and robust -0.025 0.193 -0.133
(5) AR-1 with overdisp -0.035 0.149 -0.233
(6) Unstructured with
overdisp -0.035 0.154 -0.226
Time (1) Indep with robust -0.059 0.016 -3.76
(2) Exch -0.059 0.016 -3.743
(3) Exch with overdisp -0.059 0.035 -1.704
(4) Exch with overdisp
and robust -0.059 0.035 -1.670
(5) AR-1 with overdisp -0.064 0.044 -1.438
(6) Unstructured with
overdisp -0.056 0.039 -1.226

with a log link function and Poisson errors to the total seizure count
gave a Pearson model chi-square of 614.79 for 55 degrees of free-
dom, clearly indicating some form of overdispersion. As we shall see,
this has implications for the specification of any GEE model that
we might fit as part of a longitudinal analysis of the four separate
trial counts.
Table 7.7 shows the results for treatment effect and for the trend
over visit from models with a log link, Poisson errors and additional
covariates for log-baseline and age which, for the sake of clarity, have
not been reported in the table. For illustrative purposes, the fitted
model included simple main effects and a common pattern of overdis-
persion, though Thall and Vail (1990) find some evidence in favour of
a more complicated structure. We show results from six models:
192 Design and Analysis of Clinical R i a l s

(1) an independence working model with robust standard errors,


(2) an exchangeable correlation model,
(3) as in (2) but with the scale parameter estimated by the Pear-
son chi-square/residual divided by the degrees of freedom in
order to allow for overdispersion,
(4) an exchangeable model with robust standard errors,
(5) a first-order autoregressive model allowing for overdispersion,
(6) a model with an unstructured correlation matrix allowing for
overdispersion.

Unlike the previous example, the results from the basic exchange-
able model (2) are quite different from those of the independence
working model with robust standard errors (1). However, this is not
because the exchangeable correlation matrix is grossly innappropri-
ate (though it is not especially good). Rather, as the results from
model (3) show, it is because model (2) has not allowed for overdis-
persion. It is clear that at least as much thought must be given to
the possible presence of overdispersion as to the structure of the cor-
relation matrix. The use of the robust standard errors will cope with
failures in either, and as we see from model (4), their use with the
overdispersed exchangeable model does increase the standard error
for the treatment effect beyond the value of the classical standard
error from the overdispersed model. The AR-1 Model (5) and the
unstructured or free correlation Model (6) give quite similar results.
Inspection of the table of estimated correlations from Model (6), ex-
plains why. Unlike in the respiratory disease example, the estimated
pattern is suggestive of an autoregressive rather than an exchange-
able correlation matrix.
The results from all these GEE models differ strikingly from the
results based on applying regression to the log-transformed counts.
This very much reflects the variation in the relative influence that
is attached to a small number of datapoints according to the partic-
ular mean-variance relationship assumed, a circumstance previously
illustrated using the polyp count data in Chapter 4. The analyses
Models for Non-Normal Longitudinal Data from Clinical Daab 193

Table 7.8. Estimated Correlation


Matrix for Epilepsy Data.
Unstructured

R=(.:- 1.00
0.39 0.57 1.00
)
0.24 0.30 0.41 1.00

illustrate that in general, for trials with a comparable subjects by


time structure, a greater emphasis should probably be placed in the
use of model comparisons and diagnostics to examine the issues re-
lating to dispersion in longitudinal GLMs than to those relating to
the details of the correlation structure.

7.3. RANDOM EFFECTS MODELS


The essential feature of a random effects model for longitudinal data
is that there is natural heterogeneity across individuals in their re-
gression coefficients and that this heterogeneity can be represented
by an appropriate probability distribution. Correlation among ob-
servations from one person arises from them sharing unobservable
variables. In the respiratory status data, for example, variation in
the propensity to have poor respiratory status might, in this instance,
merely reflect variation in severity of illness not captured by the sim-
ple binary response. For a random effects model, the assumptions of
the marginal generalised linear model are modified to:

g(pc.) = X'
23
.p* + d'23. ~ i ,
0 q = (b'u(p;j)
23

where p$ is the expected value of y Z j conditional on the values of


unobserved (latent) random variables q, specific to subject i; I$
is the corresponding conditional variance. The term dij is a vector
194 Design and Analysis of Clinical Trials

of explanatory variables; for example, if dlj = [l,tij],then the ele-


ments of ~i correspond to the intercept and slope of a subject specific
time-trend in the mean response. We use p* here rather than p, to
emphasise that the substantive meaning of the regression parameter
is different from that of p in a marginal model, unless the response
variable has a normal distribution. To illustrate this difference, we
will use the following simple random effects logistic regression model:
logit(Y,j = 1l.i) = PO* + P ; x i j + ~i (74
where ~i
this as:
- N ( 0 ,a 2 ) .It is convenient for our purposes here to rewrite

logit(Y,j = 1lUi) = @; + PTzij + aUi (7-3)


where Ui - N(0,l). The parameter a is a measure of the degree
of heterogeneity between subjects because the subject-specific inter-
cepts are & + aUi, i = 1,.. . ,n,and the Ui have a standard normal
distribution. In this model, the regression parameter p; must be
interpreted conditionally on the subject's own value of Ui. Zeger,
Diggle and Huang (1998) derive the marginal properties of the ran-
dom effects model in Eq. ( 7 . 3 ) ,so that the population-averaged coef-
ficient of the former can be compared with the subject-specific effect
of the latter. This requires integrating out the dependence on the
unobserved Ui; so, for example, the unconditional mean response is:

P r ( x j = 1) = 1 Pr(Kj = llu)f(u)du (7.4)

where f(.) is the standard normal density function. Using an ap-


proximation for the integral, Zeger et al. (1988) show that:
logit(Xj = 1) M (c2a2+ I)-'/'((Po* +~ ; x i j ) (7.6)
where c = 16&/(157r); consequently,
p M (c"2 + 1)- l / a p *
(7.7)
Models for Non-Normal Longitudinal Data from Clinical Trials 195

1.o

0.8
-- 0.6
II
r,
h 0.4

0.2

0.0
4 -2 0 2 4
X

Fig. 7.2. Simulation of the probability of a positive response in a random


intercept model logit Pr& = 1lUi) = -1.0 + q j + 1.5Ui where Ui is a
standard normal variable. The dot.ted line is the average over all 25 subjects
(taken with permission from Zeger, Diggle and Huang, 1998).

where c2 M 0.346. Zeger, Diggle and Huang (1998) demonstrate this


‘shrinkage’ effect by simulation. Figure 7.2 shows their results ob-
tained by assuming = -1, p; = 1 and o = 1.5. The dashed lines
show P r ( x j = 1lUi) as functions of z for each of 25 subjects, whilst
the solid line shows Pr(Y,, = 1>,calculated as the average of all 25
subject-specific functions. The solid line, which is in effect what we
would be estimating in a marginal model, is very well approximated
by a linear logistic, but with regression parameter 01 substantially
smaller that p;, as predicted by Eq. (7.7).
In practice, there is general agreement that average effects are
often of interest from a public health point of view, but may not
be so pertinent where the interest lies in scientific investigation of
the individual level process or in individual level prediction.
For a random effects extension of the GLPVI! it is possible to es-
timate parameters using traditional maximum likelihood methods.
196 Design and Analgsis of Clinical Trials

The likelihood of the data, expressed as a function of the unknown


parameters, is given by:

where a represents the parameters of the random effects distribution.


The likelihood is the integral over the unobserved random effects of
the joint distribution of the data and the random effects. Numerical
integration techniques are generally necessary to evaluate the likeli-
hood in Eq. (7.8), except in the case of linear models for normally
distributed responses.
As an illustration of the use of a random effects model, the follow-
ing logistic regression model will be fitted to the respiratory status
data:

logit(x3 = 117) = Po + /31 treatment, + ,& timezJ + ,& sex,


+ /34 age, + P5 centrez + P 6 baseline, + r, . (7.9)
Far this model, a likelihood conditional upon T for a subject is of
the form:
4

L(yiIPz,T ) = n [ P r ( y z , = lIPz, T ) ~ ~Pr(y,,


J = O ~ T ) ' - ~ ~ .J ] (7.10)
j=1

This likelihood can be averaged over some distribution for T and then
maximised over subjects to obtain the required regression coefficient
estimates. A variety of distributional forms for 7 might be assumed.
Table 7.9 shows the results in which the random effect T has been
assumed to be normally distributed. The integral calculation was
approximated by the use of %point Gaussian quadrature.
The significance of the effects as estimated by this random ef-
fects model and by the GEE models of Tables 7.7 is generally simi-
lar. However, as expected from the discussion above, the estimated
coefficients are substantially larger. Thus, while the estimated ef-
fect of treatment of a randomly sampled individual, given the set
Models for Non-Normal Longitudinal Data from Clinical Trials 197

Table 7.9. Results from Random Effects Logistic Regres-


sion Model Fitted to Respiratory Status Data.
Covariate Estimate SE a p-value
Treatment 2.14 0.572 3.74 < .001
Visit -0.12 0.125 -0.99 0.3
Sex 0.32 0.662 0.48 0.6
Age/ 100 -2.45 2.001 -1.23 0.2
Baseline 2.90 0.616 4.71 < 0.001
Centre 0.84 0.573 1.46 0.1

of observed covariates, was estimated by the marginal models to in-


crease the log-odds of being disease free by 1.3, the estimate from
the random effects model is 2.3. These are not inconsistent results
but reflect the fact that the models are estimating different param-
eters. The random effects estimator is conditional upon each pa-
tient’s random effect, a quantity that is rarely known in practice.
Were we to examine the log-odds of the average predicted probabil-
ities with and without treatment (averaged over the random effect),
this would give an estimate comparable to that estimated within the
marginal model.
In terms of covariate effects, the parameterisation of this random
effects model corresponds to the marginal model considered in Sec-
tion 7.2. The model is, however, now a model of the joint responses,
and as the tabulated predicted frequencies in Table 7.2 show, the
introduction of the random effect has captured the pattern of persis-
tence in response remarkably well.

7.4. TRANSITION MODELS


In a transition GLM, correlation amongst the observations is as-
sumed to arise from the dependence of the present observation on
one or more past values. Here we model the mean and variance of
198 Design and Analysis of Clinical Trials

yij conditional on past responses y Z j - ] ~ ,for k 2 1. The assumptions


of a cross-sectional GLM would now be replaced by:

where now pfj and are the expectation and variance of y i j con-
ditional on all x j - k for k 2 1:
An example of a simple transition model for the respiratory status
data is:

logit(Y,j = 1) = Po + p1 treatmenti + a K j - 1 j = 1,.. . ,4 (7.11)

Here the chance of respiratory status being good at a particular time


point depends on the treatment group and also on whether or not
a subject's respiratory status was good or not at the previous visit.
The parameter exp(a) is the odds ratio of good respiratory status
amongst subjects having good and poor status at the previous visit.
The parameter exp(P1) is now interpreted a s the odds ratio for good
respiratory status in the two treatment groups! for patients with
the same prior (or lag minus one) respiratory status. This odds ratio
can therefore be considered as applying to the transition rates among
states of respiratory disorder.
Table 7.10 presents a cross-tabulation by treatment group of the
respiratory status at time t , given the immediately prior status. In
the placebo group, the simple transition rates from poor to good
and from good to poor are roughly comparable, being 0.21, and 0.27
respectively. In the active treatment group, however, these two tran-
sition rates are more divergent, that from poor to good being 0.37
and that from good to poor 0.15. Thus, while the pattern is con-
sistent with the hypothesis that the active treatment might benefit
those with respiratory problems, it also suggests that the treatment
may be of no benefit to those without problems - the treatment
may not be preventative.
Models for Non-Normal Longitudinal Data from Clinical Trials 199

Table 7.10. Cross-Tabulation of Response Patterns


for Respiratory Disorder Data.
Placebo

Yt = 0 100 27 127
Column % 79.4 26.5
Yt = 1 26 75 101
Column % 20.6 73.5
Total 126 102 228
Active

Yt = 0 49 20 69
Column % 62.8 14.5
Yt = 1 29 118 147
Column % 37.2 85.5
Total 78 138 216

To investigate this possibility in more detail, we fitted a logistic


regression model which included the lagged response as an explana-
tory variable, e.g.,
logit&) = 00 + 01 treatmenti + @ 2 Y , j - l + ,& treatmenti x xj-1
(7.12)
Transition models can be fitted by maximising the conditional likeli-
hood of y Z 2 , . . . ,Y,t, given Y,l. Such models can be fitted using stan-
dard GLM software. The estimated ordinary and Huberised standard
errors are shown in Table 7.11. The latter are 10-20% larger than
the former; this suggests that the correlation in these observations
has not been fully accounted for by the inclusion of the immediately
prior response and that second and higher order terms might be re-
quired. However, the treatment by prior status interaction term is
200 Design and Analysis of Clinical Dials

Table 7.11. Results from Fitting Transition Model


Logistic Regression to Respiratory Status Data.
~ ~

Covariate Estimate Classical SE Robust SE


~~ ~

Treatment 0.822 0.321 0.380


yt-1 2.369 0.314 0.384
Interact ion -0 .O 10 0.461 0.568

Table 7.12. Results from Exchangeable Marginal


Logistic Regression Status with Effects of Lagged
(Prior) Respiratory Status.
Covariate Estimate Classical SE ztest
Treatment 1.003 0.287 3.49
Visit -0.186 0.091 -2.04
baseline status 0.973 0.301 3.24
lagged status 1.831 0.279 6.57

neither large nor at all significant, whereas the main effect of the
treatment (and prior history) is both.
The use of robust standard errors in the previous table could be
thought of as applying an IWM marginal modelling strategy to a
transition model. Equally, we could have approached the analysis by
adding the lagged response to the covariate list of a previous marginal
model. We did just that to the model of Eq. (7.1), and Table 7.12
gives the estimates for a subset of the estimated effects. Comparison
with Table 7.4 shows the estimated effect of treatment to be slightly
reduced, and the lagged respiratory status as a significant predictor,
its effect being in addition to that for the baseline response.
We have tabulated the expected frequencies for this model in
Table 7.2 for comparison with the corresponding simple marginal
model without lagged effects and the random effects model. In this
instance, as a method of modelling the joint response distribution,
Models for Non-Normal Longitudinal Data from Clinical Trials 201

there is little that has been gained by the addition of the lagged
response, particularly in comparison with the random effects model.
This is perhaps not surprising given the evidence from previous anal-
yses that an exchangeable model fits the data better than an autore-
gressive one, and that the chosen random effects model corresponds
to an assumption of an exchangeable correlation, while the transition
model corresponds rather more to the autoregressive pattern.

7.5. A COMPAFUSON OF METHODS

The way in which we have tackled the preceeding example indicates


that a mixing of the various approaches to the analysis of non-normal
longitudinal data is perfectly possible. It is, however, something
that we would suggest be used in practice only with some caution.
This stems largely from the rather different philopsophical bases of
the different modelling approaches, since as outlined in Section 7.3,
the different approaches do not attempt to estimate the same ef-
fect. Marginal models only rarely correspond to a probability model
for the process, but nonetheless are straightforward to apply and
do estimate the effects that are seemingly of immediate interest.
However, they should not be chosen uncritically, and a number of
authors have questioned the value of the parameter estimates from
such models (e.g., Lindsey and Lampert, 1997), particularly where
lagged response variables are included.
Transition and random effects models are typically put forward as
probability models. When specified correctly, their structure is one
such that they could be used to simulate data with the same prop-
erties as those of the data under study, and the parameters that are
estimated potentially (but not always nor necessarily) have a causal
interpretation. To specify such models correctly, the investigator can
and should draw upon available clinical and scientific knowledge of
the disease. However, the difficulties in achieving correct model spec-
ifications are considerable, particularly where it is wished that a full
causal interpretation be placed on all the parameter estimates. To
202 Design and Analysis of Clinical Trials

illustrate this point, consider a typical random effects model in which


the lagged response is used as a covariate. The random effect in a
typical random intercept model is included to account for between-
subjects variations in response propensity, and as such is clearly cor-
related with response, both current and past. Thus, although the
routine method for estimating random effects models assumes that
the random effect is uncorrelated with all included regressors, in
general, this is unlikely to be the case where the covariate list in-
cludes lagged responses. This sort of issue has been explored rather
more in social and economic applications of random effects mod-
els than in clinical trial settings (e.g., Chesher and Lancaster, 1983;
Pickles, 1991).
Even when pIOpeTly specified, it may be the case that. a p o p
ulation average estimate of treatment effect, rather than a fully
conditional estimate, may be required to assist in assessing results
for their implications for public health. However, such population
average estimates are easy to obtain by averaging the estimated im-
pact of treatment (on whatever scale is desired) for each subject, over
the study sample. Of course, this assumes that the study sample is
representative of the target, population, but this is anyway implicit
in the interpretation given to the parameters of marginal models as
being population average estimates.
Overall therefore, although there is currently a strong preference
for the use of marginal models, the case for them is not so overwhelm-
ing that they should be used to the exclusion of others. Indeed, in
some respects the debate as to the use of these different methods for
longitudinal analysis within clinical trials has hardly begun.

7.6. MISSING DATA AND WEIGHTED


ESTIMATING EQUATIONS

The generalised estimating-equation approach described in Sec-


tion 7.2.1 is not a full likelihood method but it is valid when data
are missing completely at random. The issue of bias of the GEE,
Models for Non-Normal Longitudinal Data from Clinical Trials 203

when the data are not MCAR, has been discussed by Kenward et
al. (1994). Robins and Rotnitzky (1995) have modified the GEE
method, allowing it to be valid under the weaker assumption that
the data are missing at random. The modification in part involves
weighting the usual GEE by the inverse of the estimated probabilities
of ‘missingness’for each subject. These probabilities can be derived
from a logistic (or probit) model of response in which observed out-
comes and/or other observed measures are covariates.

7.6.1. Missing Data in a Longitudinal Clinical


Trial with an Ordinal Response

Table 7.13 presents longitudinal data from a clinical trial in which the
response was measured on an ordinal scale of severity (0 = good, 3 =
poor) in relation to recovery from sprain injury. The principal inter-
est here lies in the impact of treatment on speeding-up the process
of recovery, i.e., a time by treatment interaction.
One relatively straightforward method of analysis is to use a pro-
portional odds model (see Chapter 4) and to tackle the repeated
measures aspect of the problem using the marginal modelling ap-
proach, most simply by assuming an independence working model
(or the essentially equivalent methods of survey research). The only
terms in the model that we will examine are those for treatment, a
linear trend for occasion and their interaction. Any correlation over
time is not being formally modelled.
However, before considering the results, an inspection of
Table 7.12 indicates that by the end of the trial, 9 out of the 30 par-
ticipants had dropped out. We compare three approaches to tack-
ling this problem of missing data. The first method uses all the
available observations and requires that we assume that the missing
observations are missing completely at random. The second method
uses ‘last observation carried forward’ to ‘fill-in’ missing observations.
204 Design and Analysis of Clinical Trials

Table 7.13. Recovery from Sprain Injury.

Treatment Group Control group


Patient ID Pain scores Patient ID Pain scores
1 211' 17 3000
2 21'0 18 11'0
3 1100 19 2220
4 110' 20 1100
5 2100 21 2210
6 21'' 22 110'
7 210' 23 2222
8 2100 24 210*
9 200' 25 2100
10 2110 26 3221
11 3221 27 22'0
12 3221 28 222'
13 3210 29 2211
14 3233 30 2211
15 332'
16 3211
* = missing value.

The third method adopts an inverse probability weighting approach


in which adjustments are made for selective loss of patients as re-
flected in differences in any measures made prior to loss. In this
method, the complete observations at times 1 and 2 receive a weight
of 1. For occasion 3, weights were obtained as the inverse of the
predicted probability from a logistic model, in which the presence
of a measure at time 3 was predicted by the time 2 pain score (lin-
ear trend term). For time 4, a similar procedure was used to give
weights that varied with the time 3 pain score. These last could
only be estimated where such a score was available and thus were
conditional upon response at time 3. Thus observations eventually
Models for Non-Normal Longitudinal Data from Clinical Trials 205

included in the analysis were: all time 1, all time 2, all available
time 3 where each was weighted by a simple weight, and all available
time 4 observations that also had time 3 observations and where each
was weighted by the product of simple weights at time 3 and time
4. In the main analysis account is taken of the fact that these non-
response weights are not frequency weights by the use of the robust
covariance estimator.
Results of the three analyses are compared in Table 7.14. The
fitted models all indicate a significant and large negative main effect
for time, reflecting a marked reduction in severity over time in the
comparison group. The negative estimate for the interaction with

Table 7.14. Longitudinal Proportional Odds Model for Pain Data


of Table 7.13.
Missing Value Log-odds 95% Confidence
Variable Method Estimate Interval
Treatment
AAO 0.688 (-0.684, 2.060)
Weighted 0.720 (-0.702, 2.142)
LOCF 0.562 (-0.806, 1.929)
Time
AAO -1.197 (-1.670, -0.723)
Weighted -1.164 (-1.668, -0.660)
LOCF -1.156 (-1.608, -0.703)
Treatment by time
AAO -0.242 (0.849, 0.365)
Weighted -0.264 (-0.911, 0.382)
LOCF -0.181 (-0.706, 0.344)
AAO = All Available Observations,
Weighted = inverse probability weighting,
LOCF = Last Observation Carried Forward.
206 Design and Analysis of Clinical Trials

treatment is consistent with treatment leading to a slightly quicker


recovery, but the effect is not significant and the confidence intervals
are wide. Thus the substantial differences in the point estimates
arising from the different treatments of missing data do not alter the
overall pattern of inference (though the unjustified narrowing of the
confidence intervals using LOCF is apparent - see Chapter 5).
The analysis undertaken in constructing the weights, in fact, cor-
responds to a model of non-response. The models examined here
made use of the immediately prior pain score a s the only predictor.
For both time 3 and time 4,the direction of effect was such that those
with more severe pain were more likely to return for treatment, but
in neither case was the association significant (p = 0.6 and 0.3: re-
spectively). The response models could have included treatment or
any other available variable. Although in many circumstances there
can be a range of different response models that yield weights that
give largely comparable results, differences can arise and the scope
for discretion in the choice of weighting model can be large. Particu-
lar care needs to be taken where any observations are associated with
unusually large weights. In t,his example, the final weights for time 4
observations ranged from 1.26 (observation 14) to 2.21 (observations
9 and 17).

7.7. SUMMARY
This chapter has outlined and illustrated the three main ap-
proaches to the analysis of non-normal longitudinal data: the
marginal approach; the random effects or mixed modelling approach;
and t,he transition model approach. Though less unified than the
methods available for normal data, these methods provide powerful
and flexible tools to a.nalyse, what until relatively recently, have been
seen as a.lmost intractable data. However, as with any such tool, care
in their use is required.
The random effects and transition models approaches, being
based on probability models, can be estimated by maximum likeli-
hood. This gives them a capability with respect to dealing with data
Models for Non-Normal Longitudinal Data from Clinical Trials 207

missing at random. Use of the weighting method described in this


chapter, however, extends the marginal approach to circumstances
where missing data may not be missing completely at random.
CHAPTER
8

Survival Analysis

8.1. INTRODUCTION
In many clinical trials, the main response variable is often the time
to the occurrence of a particular event. In a cancer study, for exam-
ple, surgery, radiation and chemotherapy might be compared with
respect to the time from randomisation and the start of therapy until
death. In this case the event of interest is the death of a patient, but
in other situations it might be the end of a period spent in remission
from a disease, relief from symptoms, or the recurrence of a partic-
ular condition. Such data are generally referred to by the generic
term survival data even when the endpoint or event being studied
is not death but something else. Questions of interest for such data
involve comparisons of survival times for different treatment groups
and the identification of prognostic factors useful for predicting sur-
vival times. Since these do not differ from the questions usually
posed about other response variables used in clinical trials, it might
be asked why survival times merit any special consideration, in par-
ticular a separate chapter? There are a number of reasons. The first
is that survival data are generally not symmetrically distributed -
they will often be positively skewed, so assuming a normal distri-
bution will not be reasonable. A more important reason for giving
survival data special treatment is the frequent occurrence in such
data of censored observations. These arise because at the comple-
tion of the study, some patients may not have reached the endpoint

208
Survival Analysis 209

of interest (death, relapse, etc.). Consequently, their exact survival


times are not known. All that is known is that the survival times
are greater than the amount of time the patient has been in the
study. This is right-censoring. A further reason for the special at-
tention paid to survival times is that measuring time is conceptually
different from measuring other quantities.
The analysis of survival data from clinical trials will be covered
relatively concisely in this chapter. Fuller accounts of survival anal-
ysis are available in Kalbfleisch and Prentice (1980), Cox and Oakes
(1984), Lee (1992) and Collett (1994). In this chapter, the notation
that we have used is the traditional one. In Chapter 9 we illustrate
the counting process notation that is growing in favour.

8.2. THE SURVIVOR FUNCTION AND


THE HAZARD FUNCTION
The analysis of a set of survival data from a clinical trial usually
begins with a numerical or graphical summary of the survival times
for individuals in the different treatment groups. Such summaries
may be of interest in their own right, or as a precursor to a more
detailed analysis of the data. Two functions describing the distribu-
tion of survival times which are of central importance in the analysis
of survival data are the survivor function and the hazard function.

8.2.1. Survivor Function


The survivor function, S ( t ) is defined as the probability that the
survival time, T , is greater than or equal to t , i.e.,
S ( t )= P r ( T > t ) (8.1)
If the random variable T has a probability density function f ( t ) , then
the survivor function is given by:

S ( t )= 4 co
f(u)du= 1

where F ( t ) is the cumulative distribution


210 Design and Analysis of Clinical Traals

Two probability distributions often used to introduce the analy-


sis of survival data are the exponential distribution and the Weibull
distribution. The probability density function of the former is:

and of the latter:

This is a Weibull distribution with scale parameter X and slope pa-


rameter y. (The exponential distribution is a special case of the
Weibull distribution with y = 1.)
Examples of each density function for a variety of parameter val-
ues are shown in Fig. 8.1
The survivor function for the exponential distribution is:

and for the Weibull distribution

Graphs of these survivor functions for various parameter values are


shown in Fig. 8.2.
Estimation of such parametric models is commonly approached
using maximum likelihood, where for individual i , the likelihood has
the form
L ~e)( = j ( t ip)dis(tip)l-di (8.7)
where di is an indicator variable for death/failure (1 = death,O =
censored) and 8 is the parameter vector of the distribution. In the
case of the simple exponential model, the ML estimate of the single
parameter X reduces to the observed sample total of failures divided
by the total follow-up time.
Survival Analysis 211

I
I
I
I
II

,, lambda-0.5
I lambda-1.0
, I
II

,
'. '. II
:' I
.. I,
..,,
%,

0 2 4 6 8 10

Time

Fig. 8.1. Weibull and exponential density functions.

Where there are no censored observations in the sample of sur-


vival times, a nonparametric survivor function can be estimated
simply as:
number of individuals with survival times 2t
S ( t )= number of individuals in the data set (8.8)
212 Design and Analysis of Clinical Trials

i;t \
I

1- lailbda-0.5
.......... lamwa=l.o
lam-2.0

I I I

0 2 4 6 8 10

Time

0 2 4 6 8 10

Time

Fig. 8.2. Survivor functions for exponential and Weibull distributions.

(This is a nonparametric estimator since no particular distribution


is assumed.)
The estimated survivor function is assumed to be constant be-
tween two adjacent death times, so a plot of S ( t ) against t is a
Survival Analysis 213

step-function. The function decreases immediately after each ob-


served survival time.
The simple method of estimating the survivor function described
by Eq. (8.8) can only be used if all the individuals are followed up
until the particular event of interest has happened to each. A number
of methods are available for estimating the survivor function for sur-
vival data containing censored observations. That most commonly
used is the Kaplan-Meier or product limit estimator. This involves
first ordering the survival times from the smallest to the largest such
that t ( l )5 t ( 2 )5 ... 5 t(n),and then applying the following formula
to obtain the required estimate:

3(t)=
jlt(,)<t
(1 - 2) (8.9)

where rj is the number of individuals at risk just before t ( j ) and


, d j is
the number who experience the event of interest at t ( j )(individuals
censored at t ( j )are included in r j ) . The variance of the Kaplan-Meier
estimator is give by:

(8.10)

We shall illustrate the use of the Kaplan-Meier estimator on the data


shown in Table 8.1. These data arise from a randomised controlled
clinical trial to compare two treatments for prostate cancer. The
treatments were a placebo and 1.0 mg of diethylstilbestrol (DES).
The treatments were administered daily by mouth and the trial was
double blind. The survival times of the 38 patients are given in
months. We are interested in determining whether there is any evi-
dence of a treatment difference in survival.
The estimated survivor functions of the two groups are shown in
Fig. 8.3. Since the distribution of survival times tends to be pos-
itively skewed, the median is usually the chosen measure of loca-
tion. Once the survivor function has been estimated, it is generally
214 Design and Analysis of Clinical l'kials

Table 8.1. Survival Times of Prostatic


Cancer Patients in Clinical Trial to
Compare Two Treatments.

ID Treatment Survival Status


1 1 65 0
2 2 61 0
3 2 60 0
4 1 58 0
5 2 51 0
6 1 51 0
7 1 14 1
8 1 43 0
9 2 16 0
10 1 52 0
11 1 59 0
12 2 55 0
13 2 68 0
14 2 51 0
15 1 2 0
16 1 67 0
17 2 66 0
18 2 66 0
19 2 28 0
20 2 50 1
21 1 69 1
22 1 67 0
23 2 65 0
24 1 24 0
25 2 45 0
26 2 64 0
27 1 61 0
28 1 26 1
Survival A nalysis 215

Table 8.1. (Continued)

ID Treatment Survival Status


29 1 42 1
30 2 57 0
31 2 70 0
32 2 5 0
33 2 54 0
34 1 36 1
35 2 70 0
36 2 67 0
37 1 23 0
38 1 62 0

Treatment: 1 = placebo, 2 = DES.


Status: 0 = censored, 1 = died.

straightforward to obtain an estimate of the median survival time.


This is the time beyond which 50% of the individuals in the popula-
tion of interest are expected to survive.
Although examining plots of estimated survivor functions is a
useful initial procedure for visually comparing the survival experi-
ence of different treatment groups in a clinical trial, we generally also
wish to test formally for a difference in survival between the groups.
In the absence of censored observations, standard nonparametric or
parametric tests might be used. When the data contain censored
observations, however, there are a number of modified parametric
and nonparametric tests that are available. Here we shall consider
only one, namely the log-rank or Mantel-Haenszel test. Essentially
this test compares the observed number of deaths occurring at each
particular time point with the number to be expected if the survival
experience of the two treatment groups was the same. A small nu-
merical example of the application of the test to hypothetical data
is described in Table 8.2.
216 Design and Analysis of Clinical Trials

Placebo
.............

0 20 40 E€l

Time in months

Fig. 8.3. Kaplan-Meier estimated survivor functions for the two treatment
groups in t.he prostate cancer trial.

For the prostate trial data we have:


~ ~~

Group 7l Observed Expected (0- E ) 2 / E


~~

Placebo 18 5 2.475 2.577


DES 20 1 3.525 1.809
Survival Analysis 217

Table 8.2. Calculation of Log-Rank Test for a Hy-


pothetical Set of Survival Times in Two Groups.
~~ ~ ~~

Survival times
~~~ ~ ~ ~

Patient Treatment Survival time Status


1 1 2.3 dead
2 1 4.8 alive
3 1 6.1 dead
4 1 15.2 dead
5 1 23.8 alive
6 2 1.6 dead
7 2 3.8 dead
8 2 14.3 alive
9 2 18.7 dead
10 2 36.3 alive
Calculating log-rank test
Time
1.6 T1 T2 Total
Dead O(0.5) l(0.5) 1
Alive 5 4 9
Total 5 5 10

2.3 T1 T2 Total
Dead l(0.55) O( 0.45) 1
Alive 4 4 8
Total 5 4 9

3.8 T1 T2 Total
Dead O(0.5) l(0.5) 1
Alive 4 3 7
Total 4 4 8
218 Design and Analysis of Clinical Dials

Table 8.2 (Continued)


~

Calculating log-rank test


Time
6.1 T1 T2 Total
Dead l(0.5) O(0.5) 1
Alive 2 3 5
Total 3 3 6
15.2 T1 T2 Total
Dead l(0.5) O(0.5) 1
Alive 1 2 3
Total 2 2 4

18.7 T1 T2 Total
Dead O(0.33) l(0.67) 1
Alive 1 1 2
Total 1 2 3
Expected number of deaths are shown in parentheses,
‘Alive’ means alive and at risk.
Observed number of deaths for T1 = 3;
Expected number of deaths for T1 = 0.5 + 0.55 + 0.5 +
+
0.5 0.5 + 0.33 = 2.89.
Observed number of deaths for T2 = 3;
Expected number of deaths for T2 = 0.5 + 0.45 + 0.5 +
+ +
0.5 0.5 0.67 = 3.11.

(3 - 2.89)2 (3 - 3.11)2
x2 = = 0.008
2.89 $- 3.11

This leads to a chi-squared value of 4.4with a single degree of free-


dom. The associated p-value is 0.0355 and there is evidence of a
difference in the survival experience of the two treatment groups.
Patients given DES appear to survive longer. The simple conceptual
Survival Analysis 219

construction of this score test has considerable appeal, one that is


valuable in communicating results to non-expert audiences.

8.2.2. The Hazard Function


In the analysis of survival data it is often of some interest to assess
which periods have the highest and which the lowest chance of death
(or whatever the event of interest happens to be), amongst those
people alive at the time. In the very old, for example, there is a high
risk of dying each year, among those entering that stage of their
life. The probability of any individual dying in their 100th year is,
however, small because so few individuals live to be 100 years old.
A suitable approach to assessing such risks is to use the hazard
function, h ( t ) ,defined as the probability that an individual experi-
ences an event (death, relapse etc.) in a small time interval s, given
that the individual has survived up to the beginning of the interval,
1.e..
h ( t ) = lim
+
Pr(event in t ,t s)
(8.11)
s-to S
The hazard function is also known as the intensity function, instan-
taneous failure rate and the age specific failure rate.
The hazard function can also be defined in terms of the cumu-
lative distribution and probability density function of the survival
times as:
h(t)= f(t) -
f (t) (8.12)
1- F ( t ) S(t)
It then follows that:
d
h ( t ) = --{lnS(t)} (8.13)
dt
and so
S ( t ) = exp{-H(t)} (8.14)
where H ( t ) , the integrated or cumulative hazard function is given by:

H ( t )= /0
t
h(u)du (8.15)
220 Design and Analysis of Clinical Trials

For the exponential distribution, the hazard function is simply A; for


the Weibull distribution, it is Xyt7-l. The Weibull can accomodate
increasing, decreasing and constant hazard functions.
The hazard function can be estimated as the proportion of in-
dividuals experiencing the event of interest in an interval per unit
time, given that they have survived to the beginning of the interval,
i.e.,

number of individuals experiencing an even in the


interval beginning at time t
& (t )=
(number of patients surving at t)(interval width)
(8.16)

The sampling variation in the estimate of the hazard function


within each interval is usually considerable. Plots of the cumulative
hazard function, obtained by summing the interval estimates over
time, are typically smoother and easier to interpret.
In practice the hazard function may increase, decrease, remain
constant, or indicate a more complicated process. The hazard func-
tion for death in human beings, for example, has the 'bath tub' shape
shown in Figure 8.4. It is relatively high immediately after birth,
declines rapidly in the early years and then remains approximately
constant before beginning to rise again during late middle age.

8.3. REGRESSION MODELS FOR


SURVIVAL DATA
The log-rank test described in Section 8.2.1 can be extended to the
case where there are either more than two patient groups to be com-
pared or where there are possible confounding factors that may be
treated as strata. But the test, and other similar tests, are limited
in their ability to fully describe and model the data. Consequently,
more complex analyses of survival data are usually performed using
some type of specialised regression model.
Survival Analysis 221

1
0 I

Fig. 8.4. Bathtub shaped hazard function.

The regression models that have been developed for survival data
are essentially of two types. The first models the hazard function
in patient groups compared to a baseline population by means of
a multiplicative model, that is to say, additive on the log-hazard
scale. The multiplicative factor is assumed to be constant over time,
in which case the model forces the hazards in the different patient
groups to be proportional, thus yielding a proportional hazards re-
gression model. Following Pet0 (1976), estimates of the hazard ratio
in the two sample case (A and B) can be obtained directly from
the Mantel-Haenzsel log-rank test statistic and its variance (see Sec-
tion 8.2.1). Alternatively, following Mantel and Haenzsel (1959), it
may be estimated as:

(8.17)

The second type of regression model commonly applied to


survival data, models the survival times directly, with covariates as-
sumed to act multiplicatively directly on the time scale, thus ac-
celerating or decelerating time to failure. The models are generally
referred to as accelerated failure tame models.
222 Design and Analysis of Clinical Trials

Within each of these two types of model, either the baseline


hazard in the proportional hazards model, or the baseline survivor
function in the acccelerated failure time model, can be assumed t o
be either fully parametric or modelled nonparametricaily. Tradjtion-
ally, however, proportional hazard models have been semi-parametric
with the baseline hazard assumed nonparametric, and accelerated
failure time models have been formulated as fully parametric.

8.4. ACCELERATED FAILURE TIME


MODELS
The accelerated failure time model is a general model for survival
data, in which covariates measured on an individual are assumed
to act multiplicatively on the time-scale, and so can be thought of
as influencing the rate at which an individual proceeds along the
time axis. Such models can be interpreted in terms of the speed of
progression of a disease. Algebraically, such models are of the form:
Si(t>= So(#zt) (8.18)
where g5z = exp(P’x) is the acceleration factor for the ith patient
compared with the baseline patient group.
The exponential and Weibull have already been introduced as
possible survival distributions. Another distribution that is fre-
quently used for survival data is the lognormal with density function:

(8.19)

where p is the mean and o2 is the variance. The log-normal distri-


bution has a relatively heavy right tail, a feature that makes it useful
for situations in which events occur later in the follow-up period.
The accelerated failure time model is most easily considered when
it is expressed in log-linear form. Letting ti denote the survival time
for the i t h subject, the model is:
In(t,) = Po + $xi + m i (8.20)
Survival Analysis 223

where o is a scale parameter and ~i assumed to have some suitable


distribution. Gaussian errors and no censoring simply correspond
to linear regression of log-survival time. More generally the model is
used with distributions such as the exponential, Weibull, log-logistic,
and gamma. It can be shown that if the errors are exponential or
Weibull, then the model is also a proportional hazards model.
Parameter estimates can be found by maximising the likelihood
function given by:

(8.21)
i=l

where 8’ = [Po,P,o] and Si = 0 if the event of interest has occurred,


and Si = 1 if the observation’s survival time is censored.
We shall illustrate the application of the accelerated failure time
model on the data from a randomised trial of chemotherapy for lung
cancer (the data are given in Prentice, 1973, and Kalbfleisch and
Prentice, 1980). The 137 patients, all but 19 of whom died during
the trial, had survival times ranging from 1 to 999 days. In ad-
dition to standard and chemotherapy treatment, covariates included
months since diagnosis, age, prior therapy, type of cancer (squamous,
small, adeno and large) and the Karnofsky Performance Index of each
patient’s functional status that ranged from complete hospitalisation
to normal self-caring.
The results under a number of different assumptions about the
distribution of the survival times are given in Table 8.3.
The log-likelihoods under the three different distributional as-
sumptions are very similar, being - 196.75 for the exponential,
-196.14 for the Weibull, and -195.22 for the log-normal. The ex-
ponential model has one less parameter than the other two, and
represents a special case of the Weibull. A likelihood ratio test of
the exponential against the Weibull gives: ~ ’ ( 1of ) 2 x (-196.14 -
196.75) = 1.22, giving little evidence in favour of a systematically in-
creasing or decreasing failure rate. The estimated Weibull parameter
224 Design and Analysis of Clinical Trials

Table 8.3. Results from Fitting Accelerated Failure


Time Model to Survival Times from the Lung Can-
cer Trial.
Exponential

Covariate Estimate SE Estimate/SE


Constant 2.59 0.75 3.45
Treatment 0.22 0.20 1.11
Months since DX -0.01 0.01 -0.03
Karnofsky Performance
Index/lO -0.31 0.05 6.00
Prior Treatment -0.00 0.02 -0.21
Cell type:
small v squamous 0.38 0.27 1.38
adeno v squamous -0.44 0.26 -1.70
large v squamous -0.74 0.29 -2.50
Log-L -196.75

Weibull

Covariate Estimate SE Estimate/SE


Constant 2.64 0.71 3.72
Treatment 0.23 0.19 1.22
Months since DX -0.00 0.00 -0.06
Karnofsky Performance
Index/lO -0.30 0.05 6.23
Prior Treatment -0.00 0.02 -0.20
Cell type:
small v squamous 0.40 0.25 1.56
adeno v squamous -0.43 0.24 -1.76
large v squamous -0.74 0.27 -2.68
Scale 0.93
Log-L -196.14
Survival Analysis 225

Table 8.3 (Continued)

Log-normal

Covariate Estimate SE Estimate/SE


Constant 1.62 0.68 2.40
Treatment 0.17 0.19 0.89
Months since DX -0.00 0.01 -0.13
Karnofsky Performance
Index/lO -0.37 0.05 7.70
Prior Treatment -0.01 0.02 -0.47
Cell type:
small v squamous -0.12 0.28 -0.42
adeno v squamous -0.73 0.27 -2.71
large v squamous -0.77 0.30 -2.59
Scale 1.06
Log-L -195.22

is very close to 1, the value for a constant hazard. The log-normal


model fits these data very slightly better than the Weibull.
Since all three models essentially fit the data equally well, it is
not surprising that they give very similar estimates for the effects
of covariates. Treatment is estimated as increasing the rate of pro-
gression along the time scale by exp(0.23) = 1.26 or 26% under the
Weibull model; and exp(0.17) = 1.19 or 19% under the log-normal
model, but confidence intervals are wide in both cases.

8.5. PROPORTIONAL HAZARDS MODEL


8.5.1. Proportional Hazards
A proportional hazards model possesses the property that differ-
ent individuals have hazard functions that are proportional to one
another, i.e., h(tlxl)/h(tlxa),the ratio of hazard functions for two
226 Design and Analysis of Clinical T ? d s

individuals with covariates xi = [ X I ~212,


, . . . ,3clp] and x; = [QI, 222,
. . . ,zzp]:does not vary with time t. This implies that, given a set of
covariates x, the hazard function can be written as:

where g(x) is a function of x and ho(t) can be regarded as a baseline


hazard function for an individual for whom g(x) = 1. The model
forces the hazard ratio between two individuals to be constant over
time since:
(8.23)

If the relative risk function g is taken a s the exponential, additive


effects on the log-linear scale are obtained. The effects of covariates
are such that the baseline hazard function, ho(t), is modified mul-
tiplicatively by covariates (including group indicators), so that the
hazard function for an individual patient is:

(8.24)

8.5.2. The Semi-parametric Proportional


Hazard or Cox Model
Although specifying a parametric form for ho ( t ) is straightforward,
such a modelling approach has been rendered largely obsolete since
Cox (1972) proposed a model and estimation method that allowed
the form of the baseline hazard t o be left unspecified. The model
is semi-parametric in the sense that only the relative risk part is
modelled parametrically.
The parameter vector P is estimated by maximising a partial
likelihood. Brief details are given in Table 8.4. Interest usually cen-
tres on the estimated regression coefficients rather than the baseline
hazard. However, an estimate of the baseline hazard function can be
obtained by a maximum likelihood approach suggested by Kalbfleisch
Survival Analysis 227

Table 8.4. Parameter Estimation in Cox’s Regression Model.

8 Assume first that there are no tied survival times and that tl < t 2 <
. . . < tl, represent the k distinct times to the event of interest among n
individual times.
0 The conditional probability that an individual with covariate vector xi
responds at time t,, given that a single response occurs a t time ti, and
given the risk set R, (indices of individuals at risk just prior to ti), is the
ratio of the hazards:
exP(P’xi)
c
jER,
exp(P’xJ

0 Multiplying these probabilities together for each of the k distinct survival


times gives the following partial likelihood function (Cox, 1975):

0 Notice that the partial likelihood is a function of P only - it does not


depend on the baseline hazard ho (t).
0 Maximisation of the partial likelihood function yields estimates of the
regression coefficients with properties similar to those of usual maximum
likelihood estimators.
0 When there are ties amongst the survival times, the likelihood function
used for estimation is usually that proposed by Breslow (1974):

where mi is the number of events at ti and si is the vector sum of the


covariates of the mi individuals with survival time ti.
228 Design and Analysis of Clinical Rials

and Prentice (1973). In the particular case where there are no tied
survival times the estimated baseline hazard function at time t ( j ) is
given by:
i L O ( t ( 2 , ) = 1- i z (8.25)

(8.26)

where x ( ~is
) the vector of explanatory variables for the individual
who experiences the event of interest at time t(q. (Since the baseline
hazard function is the hazard function for an individual having zero
values for all the explanatory variables, it is often helpful to redefine
variables by subtracting average values over all individuals in the
sample.)
The baseline survivor function can now be estimated from:

So(,) = nk

j=1
i j (8.27)

for t ( k ) 5 t 5 t ( k + l ) 7 k = 1 , 2 , . . . , T - 1. The estimated value of


the baseline survivor function is zero for t 2 t ( r )unless there are
censored survival times greater than T(T), in which case it is undefined
beyond t ( r ) . From So(t),the estimated survivor function for the ith
individual with vector of covariates xi is:

SZ(t) = [ S o ( t ) ] e x p ( B ’ x , ) (8.28)

The structure of the model as a regression model becomes easier


to see if we rewrite the model as specified in Eq. (8.24) as:

(8.29)

showing that the proportional hazards model may also be regarded


as a linear model for the logarithm of the hazard ratio. Consequently,
Survival Analysis 229

the estimated regression coefficient corresponding to a particular


covariate gives the change in the logarithm of the hazard function
produced by a unit change in the variable, conditional on the other
covariates remaining constant.
The variance of the regression coefficients derived from maxi-
mum partial likelihood are typically obtained in the conventional
way from the inverse of the negative matrix of second derivatives of
the log-partial-likelihood. But, as in the case of logistic regression,
an exponential transformation of the coefficients is often preferred,
giving more interpretable hazard ratios. However, the hazard ratio
scale typically gives a likelihood function that is not quadratic and
estimates that are not normally distributed. As a consequence, both
Wald test p-values and confidence intervals perform poorly if directly
calculated using a variancecovariance matrix on this scale. Instead,
these are generally derived from the estimates and covariance matrix
on the log-linear coefficient scale, confidence intervals being based
on exponential transformation of the corresponding end-points of
the log-linear interval.
We have seen from Chapter 4 that an alternative parameter
variancecovariance estimator is the sandwich estimator based on
the score residuals. The use of this estimator has been proposed in
the survival context by Lin and Wei (1989). For generalised linear
models (Chapter 4), this estimator is particularly valuable for clus-
tered data, a feature that is implicit within the Cox model in which
the same individual can appear repeatedly as a member of a series
of risk sets (see Table 8.4). This reoccurrence of the same individual
within the likelihood is emphasised on recognising that the likelihood
treats equivalently data deriving from one individual who contributes
to two risk sets, and data in which that individual is replaced in the
second risk set by another individual with identical covariate values.

8.5.3. Cox Model Example


To illustrate the use of Cox’s proportional hazards model, we use the
data from the lung cancer trial. For simplicity, we consider only two
230 Design and Analysis of Clinical Trials

of the explanatory variables. These are treatment and the Karnofsky


Performance Index, a continuous score that was a powerful predictor
of survival. We have previously followed others in fitting this index in
its raw form, but given the exponential inverse link of proportional
hazards models it might be more natural to fit the model to the
log-transformed index. Indeed, the model does fit slightly better
using the log-index rather than the raw index. The log-hazard ratio
estimates for this model are given in Table 8.5. Again, there is no
evidence of a treatment effect.
It is of interest to compare the width of the confidence inter-
vals from this model with those from a parametric model. Fitting
the corresponding Weibull model gave a confidence interval for the
treatment effect of width 0.694 on the log-hazard ratio scale, com-
pared to a width of 0.714 from the Cox model of Table 8.5. It is clear
that there is little loss in the precision of the estimates of interest in
using a non-parametric baseline hazard.
The estimates of Table 8.5 are for coefficients on the log-linear
scale rather than the more easily interpreted hazard ratio scale. Fol-
lowing the previous section, and taking as an example the estimated
coefficient for treatment conditional upon the Log-Performance In-
dex given in Table 8.5, the hazard rate for patients treated with
chemotherapy is exp(0.0637) = 1.066 times the hazard rate for pa-
tients given the placebo. The corresponding 95% confidence interval
is (0.75,1.52). The inclusion of the value one within this interval is
consistent with no evidence of a treatment effect (positive or nega-
tive). Variation in response to this treatment depending upon the

Table 8.5. Cox Proportional Hazards Model Estimates:


Lung Cancer Trial Data.
Log-Haz. Ratio Std. Err. z p-value
Treatment 0.064 0.182 0.35 0.7
Log-Performance
Index - 1.424 0.200 -7.13 < 0.001
Survival Analysis 231

Table 8.6. Comparison of Standard (S), Robust (R) and Weighted


(W) Estimates for Lung Cancer Trial Data.
Haz. Ratio Std. Err. z p-value 95% Conf.
Interval
Treatment
Standard 1.066 .194 0.349 0.73 (0.75, 1.52)
Robust 1.066 .203 0.334 0.74 (0.73, 1.55)
Weighted 1.077 .181 0.441 0.66 (0.77, 1.50)
Log-Performance Index
Standard 0.241 .048 -7.132 < 0.001 (0.16, 0.36)
Robust 0.241 .064 -5.358 < 0.001 (0.14, 0.41)
Weighted .312 .084 -4.304 < 0.001 (0.18, 0.53)

type of tissue affected would not be surprising. Though not shown,


it is of interest to note that the inclusion of an interaction between
treatment and cell-type shows marginal significance, consistent with
squamous types being more responsive to this chemotherapy treat-
ment than other types.
Table 8.6 presents the hazard ratios and 95% confidence inter-
vals for effects of treatment and Performance Index. In addition,
it also presents confidence intervals and p-values based on the sand-
wich estimator of the parameter covariance matrix. These are a little
larger than the corresponding values using the conventional informa-
tion matrix approach We defer discussion of the weighted estimator
shown in the table to Section 8.5.6.
The estimated baseline survivor function for the lung trial data
are shown in Table 8.7 for selected failure times. As a result of
the prior standardisation of the explanatory variables about their
mean values, this corresponds to the survivor function when all the
covariates (including treatment group) are at their mean values. We
can use the values in Table 8.7 to calculate the estimated survivor
functions for the standard and chemotherapy (68/137 = 0.496 of
the sample) treatment groups when the other covariates are at their
232 Design and Analysis of Clinical Trials

Table 8.7. Estimated Baseline Survivor Function at Selected In-


tervals for Lung Cancer Trial Data.
~ ~ ~~

Survival time Survival Prob. Survival Prob.


Standard Treatment Standard and Chemotherapy
1 1.oooo 0.9782
125 0.3532 0.2267
249 0.1261 0.1017
373 0.0540 0.0525
497 0.0094 0.0177
621 0.0095
745 0.0095
869 0.0095
993 0.0027
1117

average values as follows:

survivor function standard treatment = [S~(t)]exP~~1~066~0


(8.30)
survivor function chemotherapy = [s~(t)]exp~~1~066~1
(8.31)

Plots of the estimated survivor functions are shown in Fig. 8.5.

8.5.4. Checking the Specification of a Cox


Model
Cox’s regression model makes two key assumptions:

0 The effect of covariates is additive and linear on a log-hazard


scale.

This linearity assumption is similar to that found in most other


modelling methods we have considered in previous chapters.
Survival Analysis 233

0 400 600 800 loo0

Months surviving

Survival curve for test treatment


other covariates at mean values

0 200 400 6M) 800 loo0

Time

Fig. 8.5. Estimated survivor functions for the two treatment groups in the
lung cancer trial when other covariates are fixed at their mean values.
234 Design and Analysis of Clinical Trials

0 The ratio of the hazards of two individuals is the same at all


times.

The proportionality assumption applies to all regressors in the


model and not just the treatment effect, and a lack of proportionality
can arise for a variety of reasons:

(1) Non-instantaneous treatment benefit. A treatment may require


some time to be implemented following the randomisation time
point or may require several sessions to become effective. Some
treatments may carry short term risks which it is hoped are com-
pensated by longer term benefits. Surgical treatments typically
are of this kind, with complications leading to initial excess mor-
tality when compared to a nonsurgical treatment.
(2) The effect ‘wears off’ over time. The treatment might halt dis-
ease progression but only temporarily or the disease may become
progressively insenstitive to the treatment as the treatment pe-
riod is extended.
(3) A predictor variable may be time varying but the model repre-
sents its effect as due to a single baseline measure. As time goes
on, the baseline measure comes to reflect the contemporaneous
value of the covariate less and less well and thus becomes less
predictive of subsequent survival. For the treatment variable,
this can arise through a ‘drift’ away from the treatment proto-
col, e.g., as a result of increasing non- and poor compliance by
patients or inadequate monitoring. For covariate effects such as
measures of disease severity, individual variation in the progres-
sion of the disease will mean that severity at baseline no longer
reflects current severity.
(4) Baseline measures are measured subject to error at the time of
measurement.
(5) Effects are not uniform across patients. This might arise where
the treatment benefit applies only to a subsample of individuals,
such that over time those for whom treatment has no effect are
lost from the treated sample.
Survival Analysis 235

To an extent, the particular tests and checks that one might use
of the modelling assumptions and the extensions of the model that
might be considered, will depend upon which of these possible causes
of nonproportionality are suspected.
As in the checking of other models, residuals play a key role.
A number of different residuals have been proposed for the Cox
model.
0 Cox-Snell residual; this is defined for the ith individual as:

rci = exp(j’xi)fio(ti) (8.32)

where Z?o(ti) is the estimated cumulative baseline hazard func-


tion at time ti, the observed survival time of the individual. If
the model is correct, then rci will follow an exponential distri-
bution with mean one, regardless of the actual distributional
form of S(ti).
0 Martingale residual; this is formed by taking the difference
between the event indicator, Si and the Cox-Snell residual:

Such residuals can be used to assess whether any particular


patients are poorly predicted by the model, with large neg-
ative or positive residuals indicating a lack of fit. They can
also be used together with continuous covariates for assessing
the functional form required for the covariate with a random
scatter about zero, indicating that the variable does not need
transforming.
0 Deviance residual; this may be calculated from the martingale
residual as follows:

TDi = sign(rMi4-2irn.li + bi ln(Si - rn.li)l) (8.34)

Such residuals are particularly useful in identifying individuals


who are poorly predicted by the model, such individuals being
indicated by large negative or positive values of T D i .
236 Design and Analysis of Clinical Trials

0 Schoenfeld or Eficient Score Residuals


Schoenfeld (1982) suggested the use of residuals derived directly
from the score function of the partial likelihood. A set of partial
scores for each event can be obtained using:

"j - c
i E Rj
[Xiexp(Pz2)l

u(p)=
?j =
c
iER3
exp(Px2)
(8.35)

These compare the z vector of the subject who fails with its expected
value among all those subjects at risk. For each regressor, one (par-
tial) residual is obtained for each event, a feature as we show below
that is convenient for checking for homogeneity of effect over the
course of a trial.
Using once again the lung cancer trial data, Fig. 8.6 shows the
martingale residuals from the fitted Cox's regression model plotted

1-

31
92
0 - 81

-am
-1 -
2
!?
-mm -2-
(D

; -3- 7s
70

r 78

-5
1"' -1.5
1 I
-1 -.5
I
U

I
0
I
.5
log-performance index
Martingale residuals and log-performance index

Fig. 8.6. Martingale residuals from the Cox's regression model fitted to
the survival times from the lung cancer trial potted against the continuous
covariate.
Survival Analysis 237

1I
‘I
.....

1 7 15 24 33 42 51 60 69 78 87 96 107 119 131

Observation number

Fig. 8.7. Index plot of deviance residuals from Cox’s regression model fitted
to the survival times from the lung cancer trial.

against the continuous covariate log-Performance Index. Figure 8.7


is an index plot of the deviance residuals. The most obvious feature
of all the plots is the consistently large value associated with patient
118 (who had unusually low performance at the start of the trial,
having an index value of just lo), and to a lesser extent patient
238 Design and Analysis of Clinical Trials

o perf1 A perf2
0 perf0

6 500 1000
time
Cumulative hazard by performance stratum against time

Fig. 8.8. Cumulative hazard functions for strata defined by the Karnofsky
Performance Index from the lung cancer trial.

44 (who survived an unusually long time given their relative poor


performance index value of 40).
Numerous graphical plots have been proposed for checking the
proportionality assumption (see Chen and Wang, 1991). Such plots
should have a clear and simple pattern when the proportionality as-
sumption holds. Plots of the cumulative hazard by sub-groups, each
with a particular covariate pattern, are an obvious possibility, but
this typically requires a course grouping of patients. We grouped
patients from the lung cancer trial into three categories of the Per-
formance Index. Figure 8.8 shows the three empirical cumulative
hazard functions. These should form straight lines from the origin,
but in this case are not fully convincing.
Alternatively, a double logarithmic transformation of Eq. (8.23)
gives:
+
In[- ln(S(t))] = ,Ox In[- ln(So(t))] (8.36)
Survival Analysis 239

44

70
-1
I I I
1 500 1000
time
Scatter and smooth of score residuals for log-performance

Fig. 8.9. Schoenfeld (1982) or partial score residuals and regression smooth
plotted against log-Karnofsky Performance Index from the lung cancer trial.

Thus, plots of the empirical log cumulative hazards for subgroups


based on shared covariate patterns should, when plotted against t ,
appear parallel.
The partial score residuals introduced earlier should form a
horizontal line when plotted against time or the failure rank. The
addition of a regression smooth to the plot aids interpretation. The
slope shown in Fig. 8.9 is suggestive of nonproportionalty.
In practice, however, it is quite often difficult to assess these plots
and they can sometimes be misleading (Crowley and Storer, 1983).
Thus, in addition to graphical checks, some formal specification test-
ing is also desirable.
Cox (1972) suggested testing for time-variation in proportionality
of effects of a variable z by the incorporation of a time-dependent
variable z* = z . g ( t ) . Common choices for the function g ( t ) are
240 Design and Analysis of Clinical Trials

g ( t ) = t , g ( t ) = ln(t) and a step function such that g ( t ) = 0 if t <T


and 1 if t 2 T . We describe and illustrate how such time-varying
variables can be included in Section 8.6. A likelihood ratio test can
be performed of the models with and without the time-varying co-
variate, or the coefficient for the time-dependent variable can be
compared with its standard error for a Wald test. O’Quigley and
Pessione (1989) proposed a score test for departures from nonpro-
portionality of a similar structure. All these tests potentially involve
some arbitrary choice of the alternative, and quite often this choice
has been based on some preliminary inspection of the data, a prac-
tice which needs to be taken into account when interpreting results
(notably nominal p-values).
Gill and Schumacher (1987) suggested a test based on a compar-
ison between two hazard ratio estimators, that differ in the relative
weight that they assign to early versus late failures, in particular
the Mantel-Haenszel (see Section 8.5.1) and Prentice estimators. If
the hazards are really proportional, changes in weighting should have
little effect. For groups A and B, a hazard ratio estimator is:

(8.37)

nA .ng.
For the the MH estimator, the weights w j are given by 3
nj
3 , and
nA.ng.
for the Prentice estimator, by nj ni=,nICnk:l
-d +1
*

8.5.5. Goodness-of-fit Tests


As we have seen earlier the log-rank (MH) test is a score test that
compares the observed number of deaths in groups of patients with
those expected under a model of equal hazards. The tests of Schoen-
feld (1980) and Moreau et al. (1986) can be thought of a s an extension
Survival Analysis 241

of such an approach in which firstly, the groups to be compared repre-


sent not just a partitioning of the covariate space (patient groups) but
also of the time axis, and secondly, the expected deaths are derived
from the fitted Cox model. Crouchley and Pickles (1993) describe a
general class of specification test that is easily estimated from residu-
als, for both focussed hypotheses and for more general omnibus tests.
Although probably superior to the use of graphical methods, simu-
lations showed that the simple method for constructing the omnibus
test tended to reject too often, suggesting specification errors where
none in fact were present.

8.5.6. Influence
In the standard design of trial, the construction of the partial
likelihood means that as survival time increases, so the risk set be-
comes smaller. Thus individuals with the longest survival time con-
tribute not only through their numerous appearances in the risk sets
prior to their failure/censoring, but also to those risk sets near the
end of the trial in which few subjects are being compared. As a
result, individuals with the longest survival are most likely to be
influential points.
One simple approach to ensure that estimates are robust to the
possible presence of such individuals, is to artificially censor all ob-
servations at some earlier point in the trial. This can be considered
an extreme form of a weighted partial likelihood in which likelihood
contributions beyond the artificial censoring time are given a weight
of 0. More systematically, Sasieni (1993) suggests that the contri-
bution to the partial likelihood of each risk set be weighted by a
quantity proportional to the total number of individuals still at risk.
The Kaplan-Meier survivor function suggests itself as one such pos-
sible quantity. The inclusion of such weights requires the use of the
‘robust’ estimator of the parameter covariance matrix.
We illustrate in Table 8.6 the use of Kaplan-Meier weighting for
the lung cancer trial data. In this instance, the weights make very
242 Design and Analysis of Clinical Trials

little difference to the estimated treatment effect, but the hazard


ratio effect estimate for the log-Performance Index has moved closer
to the null value of one. The inclusion of the weights has also in-
creased the size of the standard error, reducing the corresponding
z-statistic. The results again cast some doubt on the model specifi-
cation with respect to this powerful prognostic factor.
In other respects, the influence of data points and the analysis of
influence can be approached in a similar fashion to that described in
Chapter 4 for generalised linear models. Thus subjects with covari-
ate measurements that are extreme in the space of covariate values
should be carefully checked. The score residuals described earlier
are equivalent except for standardisation to the empirical influence
function (Cain and Lange, 1984; Reid and Crepeau, 1985). Deletion
diagnostics can be pursued in the usual way (Storer and Crepeau,
1985).

8.6. TIME-VARYING COVARIATES


8.6.1. Modelling with Time-Varying
Covariates
Trials often include covariates with values that do not remain fixed
over time. It is tempting, therefore, to consider making allowance for
the changes in such variables by taking repeated measures of them
during the operation of the trial and by fitting a model that uses
these updated covariate values. In fact, this is simply done, merely
requiring the survival period of each patient to be divided up into
a sequence of shorter survival spells, each characterised by an en-
try time and an exit time, and within which covariate values remain
fixed. Thus the data for each patient is represented by a number of
shorter censored spells and possibly one spell ending in failure/death.
Datasets can be constructed in exactly this form, each record rep-
resenting one spell and containing, an entry time, an exit time, and
a censoring indicator together with the then current covariate val-
ues. With multiple spells per patient, the file thus contains multiple
records per patient.
Survival Analysis 243

Time-varying covariates are especially tractable within the Cox


model. Inspection of Table 8.4 shows how the partial likelihood in-
volves contributions for the conditional probability of failure. These
conditional probabilities involve the covariate values that apply at
the time of each failure only. The values of the covariates between
failure times do not enter the partial likelihood. All that is required
is that contemporaneous rather than baseline values of covariates be
assigned to each risk set, making the Cox model with time-varying
covariates little more complex than that for time-fixed covariates.
Indeed for covariates that change rapidly over time the Cox model is
often simpler to manage than a parametric survival model.
For illustration we divided the follow-up of the lung-cancer pa-
tients into two periods, before and after 75 days, that each contained
approximately 50% of the deaths. Patients with follow-up times
fewer than 75 days contributed just their original record. The re-
maining patients contributed one record with a censored survival
time at day 75, then a second record with their original follow-up
time and outcome but with entry into the trial given as day 75. A
dummy variable is constructed to distinguish between the two follow-
up periods. Fitting a model with an interaction between period and
treatment or period and covariate provides a test of the constant pro-
portionality of effect that the Cox model assumes. In this case, fitting
the interaction between period and log-Performance Index gave a z-
test with p = 0.035, casting some doubt on the assumption. In the
first period, the hazard ratio was estimated at 0.20, while for second
it was 0.69. This is not an uncommon pattern for a baseline measure,
the prognostic value being rather short-lived.

8.6.2. The Problem of Internal or


Endogeneous Covariates
Although simply incorporated into proportional hazard survival
models, the use of time-varying covariates in analysis of clinical trials
data should be approached with some caution. This is because their
244 Design and Analysis of Clinical Trials

inclusion runs the risk of biasing the treatment effect as a result of


their being internal, i.e., they reflect the development of the disease
process and may themselves be partly influenced by treatment. Bio-
chemical or physical measures of disease are obvious examples. This
is well illustrated in the example of Altman and DeStavola (1994).
High levels of bilirubin and low levels of albumin reflect advanced bil-
iary cirrhosis and are highly prognostic. A treatment that improves
cirrhosis will tend to reduce bilirubin levels and increase those of al-
bumin. Altman and DeStavola showed how much of the significant
and substantial estimate of treatment effect could be removed by the
inclusion into the model of updated values of either of these variables.
From the point of view of treatment effect estimation, updating these
variables is most unwise, casting unnecesary doubt on treatment dif-
ferences. From the point of view of a scientific investigation of the
development of the process and for constructing prognostic indices,
their inclusion will be of more interest.
Thus, it is important that internal or endogeneous variables
should be distinguished from external or exogeneous variables. Ex-
ternal variables are either predeterminded, e.g., a patients age, or
vary independently of survival, e.g., the weather. However, for many
time varying variables their status as internal or external is uncer-
tain, which explains our caution. It is perhaps most helpful t o think
of internal variables as being those that are ‘causally downstream’
of treatment, but the link between treatment and variable does not
have to be a direct one. Thus if the poor health of those on the worse
or placebo treatment results in their choosing to move to a more
pleasant and health promoting climate, then not even the weather
is external!

8.6.3. Treatment Waiting Times


One circumstance where the use of time-varying covariates may be
helpful is where the timing of the delivery of one or both treatments
is not under complete experimental control. Such circumstances
Survival Analysis 245

frequently arise in organ and tissue transplantation, where at the


time of randomisation, no suitably well matched donors may be avail-
able for all patients. Two comparisons then become of interest. The
first essentially defines the treatment as that given, i.e., a waiting
time of unknown duration followed by transplantation, and com-
pares survival over both waiting and post-transplant survival periods
combined. The second defines the treatment as transplantation for
which only the post-transplant survival is relevant. These correspond
to the two rather different clinical circumstances of considering the
treatment alternatives of a patient for whom a well matched donor
is already available (the second case) and a patient for whom one is
yet to be found (the first case). Without a very rigorous protocol, it
is often unreasonable to assume that the waiting time to find a well-
matched donor is independent of transplant survival, since matching
criteria are likely to be relaxed as the waiting time increases and
transplantation may only be possible if the patient is fit enough to
survive surgery.

8.7. STRATIFICATION, MATCHING A N D


CLUSTER SAMPLING
8.7.1. Stratification
We have already seen how non-proportionality may be addressed by
generalising the model by the addition of a function of the suspect
predictor variable as a time dependent covariate. However, this ap-
proach requires that the functional form describing the pattern of
time variation be specified (e.g., a linear trend). Where the variable
for which the proportionality assumption is in question is a categor-
ical variable, then a more general approach is possible using strati-
fication. Dividing the subjects into strata g, we specify a hazard of
the form:
Wlx, 9) = h o g W exp(Px) (8.38)
in which the effects for the covariate vector x (now shortened by the
omission of the stratum identifying variable g) remain assumed to
246 Design and Analysis of Clinical Trials

be proportional, but the baseline hazard is now quite separately es-


timated for each stratum. The contribution to the partial likelihood
from each failure event is modified such that the risk set over which
the denominator is calculated, is no longer all those at risk at that
time, but only the subset of those who belong to the same stratum
as the failing individual.
In practice, stratification is used to achieve a variety of goals:

(1) It can be applied to groups who either a przori may have quite
different survival patterns, e.g., for example men and women, or
post-hoc to groups defined by a variable that has failed a test of
proportionality.
(2) It can be applied so as to obtain distinct estimates of survivor,
hazard or integrated hazard functions for the groups in ques-
tion that are adjusted for the remaining covariates but which are
otherwise unconstrained. Plots of the group specific functions
can be used to check proportionality. This has particular appeal
where the grouping variable is treatment as it provides a sim-
ple graphical illustration of possible variation in treatment effect
over time.
(3) The grouping variable need not be a fixed grouping factor but
may be time varying, e.g., a variable defining the state of the
disease. Strata can also be defined to distinguish between the
different durations, e.g., first event, second event and so on in a
recurrent event process.

8.7.2. Matching
Where patients have been matched, the simplest approach to analyse
the resulting data is to specify each matched group as a separate stra-
tum. For a simpie two-treatment trial, the resulting partial likelihood
corresponds to conditional logistic regression with contributions only
from strata in which the first of the time ranked survival times in a
pair is a failure time (i.e., pairs in which either the shortest time or
both times in a pair are censored make no contribution).
Survival Analysis 247

8.7.3. Cluster Sampling


An alternative approach is to ignore the clustering in model estima-
tion but to take account of it in the estimation of the parameter co-
variance matrix by using a robust estimator. It should be emphasised
that these two approaches estimate slightly different parameters, the
first being conditional upon any effects associated with the match-
ing variables and the second being the effect marginal to/averaging
over them.
Segal and Neuhaus (1993) exploit the fact that the survival like-
lihood can be formulated as a Poisson likelihood and can thus be
made amenable to the Generalised Estimating Equations approach
to GLM estimation described in Chapter 4. The likelihood in such a
formulation for individual i in cluster j is of the form:

(8.39)

where h,(t) and H,(t) are the baseline hazard and integrated hazard,
respectively, and the Poisson rate parameter pij is specified by a log-
linear model of the form:

In pij = In H,(t) + p’x . (8.40)

The term lnH,(t) in this last equation can be an offset, specified as


ln(t) if the model is exponential or the Breslow estimate of H,(t)
if the model is a Cox model (requiring iterative updating), or it
can be estimated, for example, as a set of dummy variables defined
for segments of the follow-up time to give a piecewise exponential
model. GEE estimation of the model parameters and of the corre-
lation among individual survival times of individuals from the same
cluster j then follows in the usual way as described ‘in Chapter 6.
Yet another approach to clustered data treats the correlation in
survival times as deriving from the hazard (or log-hazard) containing
a shared random effect. This is discussed in Sections 8.10 and 9.5.
248 Design and Analysis of Clinical Dials

8.8. CENSORING AND COMPETING


RISKS
The discussion so far has given rather little consideration as to how
censored observations may have come about. Indeed, because they
seem to pose no practical difficulty for analysis, it is far too easy
to give them inadequate consideration. Clearly censored observa-
tions coincident with a data-independent end of a trial pose no spe-
cial source of concern, but this may not be the case for losses to
follow-up occurring during the process of the trial. Losses may occur
through patient noncompliance or inadequacies in the implementa-
tion of the trial. These raise all the issues and concerns relating to
missing data, discussed in previous chapters. In other circumstances,
losses may be disguised as failures, the result of some parallel sur-
vival process that cannot be suppressed during the period of the trial.
Death from other causes than the focal cause is the typical example.
On the assumption that this other process is conditionally indepen-
dent of the process of interest, then these failures from other causes
can be treated as providing censored observations from the process
of interest.
Table 8.8 compares the Cox model estimates and standard errors
for the effects of age, surgery and mismatch score on survival follow-
ing heart-transplant for an all-causes outcome and for an outcome

Table 8.8. Estimated Effects of Prognostic Indicators - Stanford


Heart Transplant Data.
Model/Effect Hazard Ratio z 95% Confidence Interval
~

Death - All Causes


log(age in years) 14.64 2.28 (1.45, 147.44)
log(mismatch score) 4.27 1.95 (0.99, 18.36)
Death - Rejection only
log(age in years) 422.78 3.24 (10.89, 16333)
log(mismatch score) 17.71 3.05 (1.79, 112.63)
Survival Analysis 249

restricted to death from rejection (all other outcomes contributing


censored observations). The data come from the much analysed Stan-
ford Heart Transplant programme (Crowley and Hu, 1997). The in-
creased estimate and significance for the effect of quality of transplant
match, when nonrejection related failures are excluded, is striking.
The potential impact of reclassifying outcomes as defined by the
censoring indicator is thus considerable and the example serves t o
underscore the need for careful prior thought being given to the mea-
surement protocol and for the importance of maintaining blindness
through to the end of trial measurement (and perhaps even beyond,
into the analysis stage).

8.9. AUXILIARY VARIABLES


Where a variable is clearly endogeneous, although we might not wish
to include it as an independent variable within our analysis, we might
be able to increase the power of our analysis by considering it as
informative about the end-point of the trial. For example, some en-
dogeneous biological marker might be used to define an intermediate
state of disease progression within what would now be a multi-state
survival model. This would have the effect of reducing the number
of fully censored observations. Treatment effects on the risk func-
tions to both the intermediate state and the final end state can then
be estimated, using the framework of competing risks described in
the previous section. Hsieh et al. (1983) and Pocock et al. (1987)
consider the analysis of two time-dependent events. However, as our
previous discussion of multiple end-points in Chapter 4 made clear,
the question of how to combine effects requires assigning some mea-
sure of relative importance to them. This has not always proved
straightforward.
An alternative is to assume that the importance of the event is
reflected in its effect on the final end-point itself. Lagakos (1977)
described an approach for using auxiliary information for this pur-
pose within a simple exponential survival framework. Finkelstein and
250 Design and Analysis of Clinical Trials

Schoenfeld (1994) present a method that uses the time to the interme-
diate state as a covariate for subsequent survival. Their simulations
suggest that although improvements in precision of the estimates of
main interest are possible, losses of efficiency are also possible where
the intermediate state is not strongly prognostic. The use of weighted
estimating equations provides one of a number of other approaches
to this problem (Robins and Rotnitsky, 1992).

8.10. FRAILTY MODELS

One of the ways in which it was suggested that non-proportionality


might come about was in a context in which individuals varied in
their risk of death in ways not reflected in the included covariates.
This lack of homogeneity has been conceptualised as implying the
presence of a random effect in the hazard function and has been
termed frailty. The sharing of frailty may also be a useful way of
considering the correlation in response of patients that have been
formed into matched groups similar on a set of prognostically relevant
variables, sampled by a multi-stage process (e.g. individuals within
families) or the correlations among repeated or multivariate survival
times that might arise from experimental response time measure-
ments. A number of authors have investigated the impact of frailty
on estimated effects (e.g., Chastang et al., 1988; Pickles and Crouch-
ley, 1995) and this can be large for endogeneous variables. This issue
also relates to the possible presence of long-term survivors or cured
individuals (Farewell, 1982). Pickles and Crouchley (1994) also con-
sidered models in which dose-response was a random effect. Such
models can now be relatively easily fitted using multilevel modelling
software (Goldstein, 1991). We illustrate in Chapter 9 the Bayesian
estimation of a frailty model for the matched pairs data.
It is very tempting to interpret the hazard function as saying
something about the development of risk within the individual. A
decreasing hazard rate is often interpreted as expressing a biological
phenomenon associated with the individual in which (s)he becomes
Survival Analysis 251

in some way stronger, more robust or less vulnerable. Unfortunately,


the hazard rate is not a pure measure of within subject variation.
High risk or frail individuals will tend to have short survival times,
resulting in a selection over time for those who were more robust
from the outset. This selection causes the usual estimated haz-
ard rate to decline, or to rise less rapidly. The inclusion of frailty
within the model can be used in an attempt to account for these
selection effects.

8.11. INTERIM ANALYSIS FOR SURVIVAL


TIME DATA
Interim analysis has been discussed in Chapter 3. The reasons given
there for its use apply equally well to trials in which the main re-
sponse variable is time-to-death, time-to-relapse, etc. Tsiatis (1981)
demonstrates a number of important results that show the distribu-
tional structure of the log-rank test, computed over time, is similar
to that for instantaneous outcome measures, allowing the theory de-
veloped for such measures to be used in the more complex survival
analysis setting. For details, readers are referred to the two papers
mentioned and the East manual (see Appendix). Here, we simply
give an example.

8.11.1. An Example of the Application of an


Interim Analysis Procedure for
Survival Data
DeMets and Lan (1994) describe the use of group sequential methods
in a randomised, double-blind, placebo-controlled trial designed to
test the effect of propranolol, a beta blocker drug, on total mortality.
In a multicentre recruitment, 3837 patients were randomised between
propranolol or placebo. The study used the log-rank test for compar-
ison of the survival patterns of the two groups and adopted the group
sequential boundaries suggested by O’Brien and Fleming. Seven
interim analyses were planned and the relevant O’Brien-Fleming
boundaries are shown in Fig. 8.10. The results of the log-rank test are
252 Design and Analysis of Clinical Tnals

Beta-Blocker Heart Attack Trial

Z**ue r

I 0
1.88
136

0
Juna
1978
May
1979
Ocl. MMcI)
1979 1980
Ocl.
1980 %Y Ocl.
1981
June
1482
Dab

Fig. 8.10. O’Brien and Fleming boundary for log-rank test in a randomised
double blind trial.

also shown as the trial progressed. On the 5th interim analysis, the
log-rank test approached but did not exceed, the critical value. On
the 6th interim analysis, the logrank statistic was 2.82 and exceeded
the critical value of 2.23. This resulted in stopping of the trial almost
a year earlier than planned.

8.12. DESCRIBING THE RESULTS OF A


SURVIVAL ANALYSIS OF A
CLINICAL TRIAL
A number of measures of effect for survival and lifetime data have
been suggested that may help communicate results to different audi-
ences. The absolute risk reduction (ARR) is given by the difference
Survival Analysis 253

in survival probabilities for the two treatments, for some suitable


choice of time t , ARR = S A ( t ) - SB(t). Using the Kaplan-Meier
estimates for S ( t ) ,a variance for the ARR can be obtained as

var(ARR) = [l - sA(t)][sA(t)]2/nA(t) -k L1 - sB(t)][sB(t)]2/nB(t)

(8.41)
where n ( t ) is the size of the risk set at time t. The ARR can be
made more concrete by multiplying by the size of the treatment arm
of the trial to give a number of deaths avoided (NDA = ARR x n ) ,
with a variance given by n times var(ARR). It is sometimes sug-
gested that the estimated number of patients in the population who
might benefit from the treatment be used in this calculation to give
an estimate of the therapeutic impact. Care is required here, in
that the size may be rather poorly known and it is in general un-
likely that the characteristics of this wider group of patients will be
like those enrolled in the trial.
An increasingly common measure is the number needed to treat
(NNT), the inverse of the absolute risk reduction (NNT = l/ARR).
Like the ARR, this must be reported together with the length of
follow-up that is being assumed. A confidence interval can be ob-
tained by inverting the corresponding endpoints of the interval for
the ARR.
Epidemiologists find the risk ratio RR a natural scale. Given by
(1 - s A ( t ) ) / ( l - S q t ) ) ,it allows effects to be described in terms of
a halving (or some other fraction) of the risk. Note that it, too,
requires specifying the relevant follow-up period. An estimate of its
variance can be obtained using Fieller’s theorem (see Marubini and
Valsecchi, 1995). A relative risk reduction or difference, 1 - RR(t),
is also sometimes used.

8.13. SUMMARY
Survival analysis is the study of the distribution of times to some
terminating event (death, relapse etc.). A distinguishing feature of
254 Design and Analysis of Clinical Trials

survival data is the presence of censored observations and this has


led to the development of a wide range of methodology for analysing
survival times. Of the available methods, the most widely used
is Cox’s proportional hazards model which allows the investigation
of the effects of multiple covariates on the hazard function. The
model has been almost universally adopted by statisticians and ap-
plied researchers particularly for survival times arising from a clini-
cal trial. But other models for survival data are also available and
one of these, the accelerated failure time model, is also described in
this chapter.
Whatever the particular form of survival model chosen, a num-
ber of common issues arise, principally in the area of model checking.
Although less well developed than for continuous response measures,
various methods for model checking are available. These include
model generalisations, residuals, plots and goodness-of-fit and spec-
ification tests. However, in a number of instances these methods do
not work well as a reliable diagnostic for overall correct model spec-
ification. Instead, testing of survival models works best when the
statistician has some idea of the likely alternative models. In such
circumstances, it is understandable why statisticians so frequently
resort to the use of the Cox survival model, since its nonparamet-
ric baseline hazard leaves one less aspect of the model to check, and
little loss of efficiency is incurred when compared to parametric mod-
els. Nonproportionality can often be addressed by the careful use
of strata, time-varying variables are readily included, and exten-
sions to competing risks by redefinition of the censoring indicator
is straightforward.
CHAPTER9

Bayesian Methods

9.1. INTRODUCTION
Until relatively recently, Bayesian statistics were Iittle more than an
intellectual curiousity; rich in conceptual insight but of little practi-
cal value when it came to actual data analysis. All this has changed
in the most dramatic fashion, with Bayesian methods and appli-
cations now forming an area. of the most intense activity. In many
cases, Bayesian approaches lead to the same or similar conclusions as
those using the routine procedures of the frequentist statistician. But
differences do occur and there are proponents on each side with, for
example, Berry (1993) presenting the case for greater use of Bayesian
methods in clinica.1trials and Whitehead (1993) presenting the case
for the ongoing dominance of the frequentist approach, at least. in
the context of definitive phase I11 trials. Most statisticians involved
in trials are typically rather pragmatic, for the most part using the
familiar and quick to apply frequentist methods, but nonetheless also
applying Bayesian methods where these are convenient or have some
conceptual advantage and are acceptable to the intended consumer.
Much recent work proposing Bayesian methods in clinical trials has
been concerned a s much with establishing acceptability as with de-
veloping the methods.
Traditional frequentist analysis essentially treats each trial or
experiment as if it were entirely novel and each t.rial is usually be-
ing considered as being individually potentially decisive. Scientific

255
256 Design and Analysis of Clinical Trials

progress may occur outside the narrow focus of this trial, but the
numerical procedures themselves are not formulated to reflect the
process of progressive learning nor one in which the process itself in-
volves costs and potential benefits. By contrast, the focus of Bayesian
methods is one of progressive refinement of opinion as data from tri-
als and other sources accumulate. This is illustrated in Fig. 9.1.
Knowledge prior to a trial is synthesised and formally represented
as a distribution over the parameter space of the problem. The trial
is undertaken. The data from the trial is combined with the prior
distribution to form a posterior distribution over the same parameter
space, one which is hopefully more concentrated than the prior.

I-
prlor

I A.
Dlstrlbution

Trertmant
Dlftcrence

Bayes'
[Data1 I
0
T Mean
*
Theorem

Posterior
I
Distribution
Olfterence

I
(Decisions]
0

Fig. 9.1. Conceptual framework for Bayesian analysis.


Bayesian Methods 257

While this may describe well the circumstances of the developers


of a treatment, who will be building up knowledge about a treatment
through various phases of the development and trialling process, li-
censing authorities have typically been comfortable with the frequen-
tist position, the results from a phase-I11 trial, complete with all its
formal rigours, being key to their decision making. Nonetheless, even
in the context of phase111 trials, Bayesian methods can provide valu-
able additional insights, particularly in relation to considerations of
sample size requirements and interim analyses.
As described above, the focus of Bayesian methods is on using
data to update prior beliefs, as defined by parameter or effect distri-
butions. If one starts from a position of diffuse and uninformative
prior distributions, then the focus of the analysis is little different
from that of other methods. As a result, non-Bayesian’s are in-
creasingly making use of the recent developments in Bayesian model
estimation. We therefore start with a brief description of Bayesian
estimation, before considering other more distinctive aspects of
Bayesian inference for clinical trials.

9.2. BAYESIAN ESTIMATION


From a Bayesian perspective, both observed data and parameters
are considered as random quantities. Letting D denote the observed
data and 0 the model parameters, a joint probability distribution
or full probability model P ( D ,0 )is considered, which is decomposed
into a prior distribution for the parameters P ( 0 ) and a likelihood
P ( D ( 0 )for which:

P(D,0 ) = P(DIO)P(O) (9.1)


Given data from a trial, the posterior distribution of the parameters
0, that is the distribution of 0 given the data, is obtained by use of
Bayes’ Theorem,

/
P(0ID)= P(O)P(D[O)/ P(O)P(D[O)d0 (9.2)
258 Design and Analysis of Clinical Trials

Quantities calculated from this posterior distribution of the


parameters form the basis of inference. Point estimates might be ob-
tained by calculating the mean, mode or median of a parameter pos-
terior distribution, while the parameter precision can be estimated
by the standard deviation or some suitable inter-quantile range, e.g.,
from quantiles at p and 1 - p for a l O O ( 1 - 2p)% credible interval.
In general such quantities, f(@), will be estimated by their posterior
expectation given by:

E [ f ( O ) l D ]= / /
f ( O ) P ( O ) P ( D I O ) d @ / P(O)P(DI@)d@ (9.3)

As an illustration of the mathematics of Bayesian inference, con-


sider a sequence of Bernoulli trials. At the outset, we will assume
that we can characterise any prior knowledge that we have as to the
likely value of the Bernoulli probability 8 in a prior distribution. It
is mathematically convenient to choose a so-called conjugate distri-
bution as the prior distribution for 8, which in this case is a beta
distribution with density b(p):

~ ( 8=) r l ( i - o)@-lr((Y +
p)pya)r(p) (9.4)
+
This has mean (Y/((Y p) and variance a p / ( ( a ,@'(a ?!, + + +
1)).
Figure 9.2 shows four beta distributions. Symmetrical unimodal dis-
tributions are obtained for (Y = p > 1, narrowing as their value
increases, becoming asymmetrical when a does not equal p. Also
notice that although a = p = 1 is uniform over 8, calculation of
the variance shows that larger variances are obtained from bimodal
shapes in which (Y and p tend to zero.
In a sequence of n trials we observe T successes, giving a like-
lihood P ( n , r ( B )proportional to O'(1 - Q)n-'. Equation 9.2 shows
that the posterior distribution is obtained by multiplying the prior
distribution by this likelihood and standardising. The selection of a
conjugate prior makes this straightforward, giving

P(8ld, n ) 0: Oa+'-l(1 - qP-tn-r-1 (9.5)


Bayesian Methods 259

0 1
0

Fig. 9.2. Four beta distributions.

which is another beta distribution but with parameters a* = a I-+


+
and p* = p n - I-.
+
The mean of this posterior distribution is ( a .)/(a + +p n),
with a corresponding expression to that given above for the vari-
ance. As a consequence as a and p approach zero, corresponding
to a prior distribution with the greatest possible variance, so the
posterior mean approaches r / n , the value that would be expected
under maximum likelihood. As T and n increase relative to a and
p, so the the variance of the distribution approaches the familiar
(I-ln)x (1 - ( r / n > > / n .
Figure 9.2 shows a beta distribution with a = p = 6.5, the
posterior distribution that would occur following the observation of
6 successes in 12 trials with the use of the reasonably uninformative
prior in which a = p = 0.5 (also shown).

9.3. MARKOV CHAIN MONTE CARL0


(MCMC)
In the hypothetical example of the previous section, the particular
choice of likelihood and prior allowed the necessary expectations to
260 Design and Analysis of Clinical Dials

be calculated by hand. In practice, this is not generally possible.


Instead Murkov Chain Monte Carlo methods are used; these are
based on Monte-Carlo integration methods for tackling the numerical
problems posed by Eq. 9.3. We give here only the briefest sketch of
MCMC estimation.
Monte Carlo integration consists of a method for drawing values
of 0 from the posterior distribution P ( Q ( D ) ,calculating the corre-
sponding values f ( @ ) and then using their average t o approximate
E [ f ( O l D ) ]Thus
. for a ‘sample’ of m such values of 0 :

For many methods of sampling from P(QlD), this approximation


can be made as accurate as required simpy by increasing the size of
the ‘sample’ m. One of these methods is to sample from a suitable
Markov Chain that has as its stationary distribution P ( O ) D ) ,giving
rise to the name of the overall estimation method.
One of the features of such a procedure is that, regardless of where
it is started, in the long-run the state occupancy distribution tends
to converge t o the equilibrium stationary distribution. Thus the
MCMC method consists of a ‘burn-in’ during which it is intended
that the stationa.ry distribution should be achieved, followed by a
period of ‘monitoring’ during which ‘sample’ values of the quantities
of interest f(Q)are recorded and during which tests of convergence
are undertaken (Gelman, 1996). Readers are referred elsewhere for
the theoretical background and for further discussion of available
sampling algorithms (e.g. Roberts, 1996).

9.4. ADVANTAGES OF THE MCMC


APPROACH
Although the principles behind the MChlC estimation and inference
are no more complex than those of more traditional methods, it is
a computational intensive method to put into practice. Therefore,
it is helpful to have clear what some of the potential advantages are
before setting out to implement such an analysis.
Bayesian Methods 261

9.4.1. Model Parametrisation and Scientific


Inference
The effects reported its the results of any model estimation exer-
cise, e.g., maximum likelihood, are typically presented in the form
in which the statistical model has been parametrised for estimation.
Only sometimes are estimates and confidence intervals translat,ed
into a more readily understood scale. An example of this is logistic
regression where considerations as to the multivariate normality of
the estimators leads t o estimation of effects on the log-odds scale,
even though for most clinical audiences it is the odds-ratio scale that
is most readily interpreted. As a result, it is perhaps more often
the case than most statisticians appreciate that the parameters and
associated confidence intervals commonly estimated are not those of
greatest scientific interest or clinical relevance. The parametrisation
with desirable statistical properties is rarely that with desirable sci-
entific properties and the effort in translating results from one to the
other, for example using the delta method, often holds little appeal to
the statistician. However, in MCMC estimation, the parametrisation
0 over which the Alarkov Chain is defined, does not constrain the list
of quantities f ( @ ) for which posterior distributions are monitored.
Thus 0 can be chosen for its statistical and estimation properties,
whiIe the f(@> can be chosen for their scientific and clinical interest.
The additional burden of adding into the MCMC sampling cycle a
variety of functions of the parameters is rarely great.
As an example, consider a study involving three drug treatments
A, B and C and a control treatment; a standard parametrisation
would be a mean contrast for the effects of each drug relative to the
common control group. However, we might wish to know what the
probability is that the pair of drugs that perform best in some small
trial actually contains the ‘best’ drug. This kind of information is
extremely valuable in drug development, but is not readily calculated
from knowledge of the point estimates and covariance matrix of the
standard parameters. It is, however, an extremely simple task to
262 Design and Analysis of Clinical K a l s

monitor the ra.nks of the effects of each treatment from each MCMC
sample, and to obtain an estimate of their distribution, confidence
intervals and so on.
A second example, one that we illustrate later in this chapter, is
the estimation of the Number Needed to Treat (NNT) effect measure
together with its confidence interval. A variety of other effect scales
might be contemplated and the discussion in Chapter 5 relating to
the summary statistic approach is relevant here, both for pointing to
aspects of the trial that may be of interest, but also in warning of
the need to have agreed a single decision criterion prior to the study,
whenever that is appropriate.

9.4.2. Missing Data


No new ideas or methods are required to extend MCMC estimation
to problems with missing data. The missing observations are merely
added to the parameter list, and like the original parameters, are
sampled from their conditional distributions. Monitoring of the esti-
mated missing data values can be helpful in assessing the plausability
of a model. The approach can be extended to consider ‘hypothetical
data points’, for example those in the future, allowing an exploration
of an individual subject’s prognosis (Berzuni, 1996)

9.4.3. Prior Distributions as a Numerical


Device
Although frequentists, in particular, may feel uncomfortable about
the imposition of informative priors on parameters, there are many
circumstances where the imposition of such priors merely act to keep
parameter estimates within feasible bounds. In more standard ML
estimation, such problems often require the use of a variety of some-
what ad-hoe ‘fix-ups’ that substantially complicate function rnaxirni-
sation algorithms and present additional problems where convergence
is at a boundary solution. Imposing a suitable prior can result in a
better behaved estimation procedure.
Bayesian Methods 263

9.4.4. Alternative Models and Model Selection


A Bayesian approach has often also been distinctive in showing a
willingness to consider a range of alternative models, and, for exam-
ple, using so-called Bayes factors to choose between them. These
factors provide a summary of the evidence given by the data D in
favour of a model M I relative to another model Mo; they are defined
as the ratio of the posterior to prior odds,

(9.7)

Twice the logarithm of Blo is on the same scale as the deviance and
the likelihood ratio test statistic. The following scale is often useful
for interpreting values of Blo:

21nBlo Evidence for MI

<O Negative (supports Mo)


0-2.2 Not worth more than a bare mention
2.2-6 Positive
6-10 Strong
> 10 Very strong

Sometimes ‘robust’ estimates are considered that attempt to


make a combined inference from a weighted average of the parameter
estimates from different models. Considering a range of alternative
models is clearly valuable, but we are less convinced about the value
of this combining of different estimates. For example, when consider-
ing models with and without a nuisance parameter (say the effects of
a possible confounder), it seems unreasonable to consider the prob-
lem as a bimodal one - a mixture of null and alternative values -
rather than using prior information on what the value of the nuisance
parameter is likely to be.
264 Design and Analysis of Clinical Trials

9.5. A NUMERICAL EXAMPLE OF USING


BAYESIAN ESTIMATION
To illustrate the use of a Bayesian estimation method for an essen-
tially non-Bayesian analysis, we have turned to the classic data of
Frierich et al. (1963) on 6MP and placebo treatment of 42 leukemia
patients. These data are shown in Table 9.1. Though much analysed,
the paired design of this study has typically been ignored. Spiegel-
halter et al. (1996) use the program BUGS (Gilks et al. 1994; see
Appendix) to estimate a Cox regression model (see Chapter 8) that
includes a random effect for pairing.

Table 9.1. Pairwise-Matched Trial of Remissions in Acute


Leukemia (Fkierich et al., 1963).
Survival Time Death Treatment Pair ID
(weeks)
(+0.5 = 6-mercaptopurine)
1 1 0.5 1
1 1 0.5 2
2 1 0.5 3
2 1 0.5 4
3 1 0.5 5
4 1 0.5 6
4 1 0.5 7
5 1 0.5 8
5 1 0.5 9
8 1 0.5 10
8 1 0.5 11
8 1 0.5 12
8 1 0.5 13
11 1 0.5 14
11 1 0.5 15
12 1 0.5 16
12 1 0.5 17
15 1 0.5 18
Bayesian Methods 265

Table 9.1. ( Continued )


Subject ID Death Treatment Pair ID
17 1 0.5 19
22 1 0.5 20
23 1 0.5 21
6 1 -0.5 19
6 1 -0.5 18
6 1 -0.5 8
6 0 -0.5 1
7 1 -0.5 20
9 0 -0.5 6
10 1 -0.5 2
10 0 -0.5 10
11 0 -0.5 3
13 1 -0.5 14
16 1 -0.5 4
17 0 -0.5 11
19 0 -0.5 7
20 0 -0.5 9
22 1 -0.5 12
23 1 -0.5 16
25 0 -0.5 17
32 0 -0.5 5
32 0 -0.5 13
34 0 -0.5 15
35 0 -0.5 21

Using the counting process notation (Anderson and Gill, 1982),


we observe the count of the number of failures up to time t , Ni(t)
which has the intensity process:

I&)& = E[dNz(t)lFt-] (9.8)


where Ft- represents the data on the process up to time t. This
expectation corresponds to the probability of subject i failing in the
266 Design and Analysis of Clinical Wals

+
interval [t,t d t ) . As dt -+ 0 and assuming a proportional hazards
model with time constant covariates and time constant random effect
for each pair, the intensity takes the form:

where j ( i ) is the pair to which patient i belongs. In line with many


of the random effects models described in Chapters 6 and 7, the
distribution of the random effects can be assumed normal, implying
in this case a log-normally distributed multiplicative effect on the
intensity itself.
The parameters of the model, therefore, include the regression
parameters 0, the baseline intensity A,@), and the precision (l/vari-
ance) of the random effect distribution. Spiegelhalter et al. choose a
normal prior with very low precision as an uninformative prior for p.
For the baseline intensity, it is convenient t o consider the increments
between failures as having a log-normal prior. For the precision of
the random effect distribution, a gamma prior is convenient.
Having specified the prior distributions and starting values for
the parameters, parameter estimates can be obtained through the
use of Gibbs sampling. This involves drawing values from the con-
ditional distributions of each parameter, given the data and current
values of the other parameters, taking each parameter in turn. This
process can be shown to converge, such that after a suitable ‘burn-in’
period, typically of the order of several and often many hundreds of
iterations, the distribution of the parameter vector remains the same
from iteration to iteration.
At this point, the variation in the values of the parameters from
iteration to iteration represents the variance (and covariance) in
their estimates as a result of having a limited amount of data. With
uninformative priors and a well behaved problem, this typically cor-
responds closely to the sample variancecovariance matrix of parame-
ter estimates that might be obtained by direct maximum-likelihood.
So, after the ‘burn-in’ period, the stream of sample values for the
Bayesian Methods 267

Table 9.2. Comparison of Cox Model Estimates for Leukemia


Trial Data.
Partial Likelihood MCMC MCMC
SAS PHREG Without With
Pairwise Pairwise
Frailty Frailty
Treatment effect 1.59 (0.43) 1.55 (0.43) 1.58 (0.43)
Random effect SD 0.16 (0.03-0.69)

parameters are monitored and stored, and the empirical mean,


medians, variances: covariances, confidence-intervals and so on are
estimated directly from these values. Clearly, the number of Sam-
ple values upon which these are estimated must be large enough to
keep the simulation sampling error small. For the calculation of some
statistics, such as exbreme confidence intervals based on the empirical
distribution (rather than a variance estimate), the simulation sample
may need to be large (in excess of IOOO}.
Table 9.2 presents results of standard maximum-likelihood fitting
of this Cox model without a frailty effect, and those obtained using
Gibbs sampling with and without the frailty random effect. The
variation in the estimates of the fixed treatment effect have rather
more to do with the assumptions being made within each method
about the treatment of ties (see Chapter 8) than with the particular
estimation method, but this variation is anyway small. The estimate
of the precision of the random effects distribution is consistent with
a random effect variance of only 0.03, suggesting that the pairings
were not well matched or were matched on the basis of factors and
covariates of little prognostic significance.
As described above, within the iterations of the sampling algo-
rithm additional functions of the parameters and data can be calcu-
lated. We added a function to estimate NNT. The posterior mean
for NNT was estimated as 2.51, with a standard deviation of 0.96
and 95% credible interval of (1.46,4.87).
268 Design and Analysis of Clinical Trials

9.6. INFORMATIVE PRIORS

In the previous example the priors were essentially uninformative,


chosen largely to keep parameters within a feasible space, and
their presence within the analysis had very little influence on the
eventual results. If, however, informative priors are to be considered,
the statistician enters what is for many the rather unfamiliar terri-
tory relating to the characterisation of prior beliefs. It is often con-
venient, both conceptually and in practice, to consider this process
as one of translating prior beliefs into a hypothetical dataset. This
dataset then informs a design in the same way that a meta-analysis
of previous findings can inform a proposed study (see Chapter 10).
The choice of information source for a prior depends on the pur-
pose to which it is to be put. For example, where studies are being
undertaken within the context of an ongoing research programme,
it makes sense to analyse and interpret the results of new studies
as part of an ongoing accumulation of knowledge. However, while
use of previous trials is appropriate in the assessment of final results,
Freedman, Spiegelhalter and Parmar (1994) suggest that such data
should play a more minor role in monitoring a new trial if it is to pro-
vide independent confirmation. Freedman and Spiegelhalter (1983)
describe methods which we describe in the next section for elicit-
ing subjective opinion. Stewart and Parmar (1993) discuss issues in
relation to using previous studies.

9.7. PRIORS AND STOPPING RULES FOR


TRIAL MONITORING

We have already described in Chapter 3, how for many large trials,


trial monitoring and the use of interim analyses are now routine
practice, required both for ethical and financial reasons. It was also
noted how the Bayesian approach to sequential analysis helped avoid
a number of the difficulties encountered by the frequentist approach
to this problem. It therefore makes sense to develop the Bayesian
argument within such a context.
Bayesian Methods 269

- Confidence interval Decision

Stop and recommend standard treatment

- Continue trial

Stop and recommend new treatment

Use standsrd meatmeut


-- -1
Usenew meatment
>5
6, 6,
Range
of
equivalence

Fig. 9.3. Ranges of equivalence elicited from clinicians.

Freedman and Spiegelhalter (1983) proposed a framework for


considering a new treatment versus a standard treatment. Two as-
pects of clinical opinion need to be considered: clinical demands and
clinical beliefs.

9.7.1. Clinical Demands


Freedman and Spiegelhalter suggested that two levels of treatment
improvement should be considered, as illustrated in Fig. 9.3. The
lower value SL is the treatment effect below which a clinician would
definitely not use the new treatment. It is referred to as the mini-
mum clinically worthwhile difference. In this hypothetical case it is
slightly positive, perhaps the result of the new treatment requiring
new training, extra costs or patient discomfort. The upper va.lue 6u
is the level of improvement above which the clinician would wish to
switch to using the new treatment as routine. The range between 6~
and Su is referred to as the range of equivalence. As a study pro-
ceeds, decisions on continuing the trial are then made by comparing
270 Design and Analysis of Clinical Trials

the confidence interval for the treatment effect with this range of
equivalence. Where the confidence interval falls entirely above SL,
the study should be stopped and the new treatment adopted as rou-
tine. Where the interval falls entirely below Su,the study should be
stopped and the standard treatment continued as routine. If neither
applies, then the trial should continue.
The values of SL and Su were obtained by interview with a range
of clinicians involved in the treatment of the disease in question.
Clinicians responded to the questions of the sort, “if the real ben-
efit was x%, would you use the new treatment as your routine?”
Such questions were asked a number of times, changing the value
of x up and down the scale, to identify the levels at which the
clinician would definitely use (Su)and not use (SL) the new treat-
ment. Such questioning yields results as illustrated in Fig. 9.4, with
both the location and the width of the range of equivalence varying
from clinician to clinician. The extent of overlap between the prior
distribution and the range of equivalence provides the ethical basis
for randomisation. Freedman and Spiegelhalter found considerable

2 5 1
0
-
C
u
u
W
.-I

I
m
U
U -

0- 11
I I I
Use standard treatment

I I
Bayesian Methods 271

variation between clinicians in the elicited ranges of equivalence. This


raised doubts as to the applicability of the simple decision rules de-
scribed in the previous paragraph.

9.7.2. Clinical Beliefs


In eliciting views on clinical beliefs clinicians were first asked what
the ‘most likely’ level of improvement would be, then upper and
lower bounds that were ‘very unlikely to be exceeded’ together with
a judgement of what odds corresponded to ’very unlikely’, and then
similar questions for intermediate points. A sketch of this ’prior’
was then verified and adjusted through discussion. There are, of
course, a variety of other ways in which such distributions might be
elicited. Parmar et al. (1994) present an example that makes use of
analogue scales and the allocation of % beliefs over a set of ordinal
categories. As in the case of clinical demands, substantial diversity
of opinion was found, with marked bebween-clinician variation in
location, dispersion and shape. As we will see: this heterogeneity of
opinion, together with the fact that clinicians involved in a trial from
whom priors are most likely to be sought and received, are likely to be
enthusiasts for the new treatment, has prompted the consideration
of a different approach to trial monitoring that we describe in a
later section.
A summary of prior opinion can be formed from these individual
assessments by approximating them by a fitted distribution. For
surviva.1 data, a convenient scale on which to consider this issue is
the log-hazard ratio, LHR, for which:

and P,,, and P s t a n d a d are the survival rates (proportions) under


the new and standard treatments, respectively. Assuming that the
distribution on the log-hazard ratio is normally distributed, the prior
can be characterised by a mean po and variance a:. Since for survival
data (the Cox model) the variance is approximately equa.1 to 4 / n ,
2 72 Design and Analysis of Clinical Traals

where n is the number of events (Tsiatis, 198l), the prior can be


summarised by a mean po and an implicit number of events no.

9.8. MONITORING
9.8.1. The Prior as a Handicap
Subsequent real data arising out of the trial with which to update the
prior are summarised by corresponding values pd and nd. Combining
the prior and data together using Bayes theorem gives a posterior
+
mean p p , given by the simple weighted sum (po x no pa x n,-J/(ng +
+
nd) with variance 0; given by 4/(no n d ) . Undertaken early in the
trial with little real data, the combined results will largely reflect
the prior. Thus only the most extreme data could result in confidence
intervals that did not include one or the other of the SL and Su,
with correspondingly little chance of very early stopping of the trial.
Undertaken late in the trial, t,he weight of data will dominate the
influence of the prior, generating results that are similar to those
of a traditional analysis. A typical prior thus acts as a handicap,
restraining the study in the early phases from premature conclusions.
Grossman et al. (1994) provide a more structured approach for
how the handicap can be used for interim analyses within a group se-
quential design. They consider a trial involving T sequential blocks,
each of n/T patients, with analyses folIowing the completion of each
block. In addition, a handicapping prior sample of size f x n,the
size of f reflecting the extent of handicap desired, is included. Thus
+
at the tth analysis, the sample size is f n nt/T. If each block of
patients has mean response yt, and the standard deviation of the
response is 0 ,then the stopping rule test statistic is 2, = (Yl + Y2 +
. . .+Y~)-\/(n/t~)/c. Without adjustment, the trial would stop if the
test statistic exceeded the usual critical value Z,, 1.96 in the case of
a = 0.05. Grossman et al. (1994) show how the inclusion of the
handicap sample in a trial with t analyses increases the critical value
+
of the test statistic by the factor J ( ( t f T ) / t . Freedman, Spiegel-
halter and Parmar (1994) compare this 'unified approach' to trial
Bayesian Methods 273

monitoring to the stopping rules of Pocock and of O’Brien and


Fleming, and suggest the following handicap as fixing the Type I
error rate at 5% for the specified number of analyses (given in brack-
ets): 0.00 ( l ) , 0.16 (a), 0.22 (3), 0.25 (4), 0.27 (5), 0.29 (6), 0.30 (7),
0.32 (8), 0.33 (9), and 0.33 (10).

9.8.2. Recognising Heterogeneity in Priors:


Enthusiasts and Sceptics
The diversity of clinical opinion identified earlier suggests that
searching for a decision rule based on the beliefs of some average or
representative view may not be appropriate. An alternative basis for
monitoring emphasises how different groups of clinicians interpret the
results. This approach addresses the question of whether a typical
sceptical clinician will be persuaded of the positive value of the new
treatement and the complementary question as to whether a typical
enthusiast will be persuaded of its inneffectiveness (see Spiegelhalter,
Freedman and Parmar, 1994). A study should be aiming to do one or
other of these. The prior of a typical enthusiast can be elicited from
the participating clinicians using the techniques described above; it
already being noted that involvement in a trial tends to select for
clinical enthusiasm for the new treatment. The suggested prior of a
typical sceptic is based on the experience in most fields that most
new treatments have been found to be inneffective or at best only
slightly better than the standard treatment. Such a prior is one cen-
tred on 0 with only 5% above the point of the alternative hypothesis
(PO),i.e., a standard deviation 00 equal to the alternative hypothesis
value divided by 1.645. This alternative value should be ‘realistic’
(not that often optimistically large value that is so often used in
power calculations to justify the small size of a trial!). As with other
priors, this can be conveniently thought of as a hypothetical dataset
from a trial, in this case of NOpatients, where NOis given by 2a/ao
with 00 as previously defined and o the standard deviation of the
response measure. The enthusiasts prior is then centred on S with
o = oo.
274 Design and Analysis of Clinical Trials

Now, the usual sample size formula gives:

N = 4[z(l - a ) +~ (-1@)I2 x a2/h2 (9.11)

where a is the one-sided significance level, 1 - ,B is the power, 6 is


the alternative hypothesis and CT is the standard deviation of the
response. But given that CTQ = 6/z(l - y), where y is the prior
probability of exceeding 6, then:

002 = 4[z(1 - a ) + z(1 - p)I2 x (Ti/” * z(1 - 7)2] (9.12)

Combining these equations and substituting 2a/No for (TO, gives:

No/N = z(1 - ,y)2/[z(1 - a ) + 4 1 - P>I2 (9.13)

where No/N is the ‘handicap’ of the previous section.


To illustrate the use of the enthusiast and sceptical priors, we
consider a hypothetical trial with failure events as the response end-
point. A new and a standard treatment are to be compared by
means of the logarithm of the hazard ratio, with the enthusiasts
expecting to see a hazard ratio of 2 in favour of the new treat-
ment. The enthusiasts’ prior is assumed normal and is centred on
ln(2) = 0.693. The sceptics’ prior is normal and centred on zero
(corresponding to an expected hazard ratio of 1). With common
variance equal to (0.693/1.645)2 or 0.176, the sceptics expect an ef-
fect size as large as the enthusiasts’ mean only 5% of the time. The
sceptics’ prior can be considered as corresponding to a hypotheti-
cal dataset in which the hazard ratio is 1 and there have been 22.7
+
failures ((nl n2) = 4/0.176).
At an interim analysis, there are 30 events under the old treat-
ment and 10 under the new, with an estimated hazard ratio of 2.9.
These data can be represented by a normal distribution centred on
+
log(2.9) = 1.065, with variance (1/30) (1/10) = 0.133.
The enthusiasts are jubilant. Under their prior, the posterior is
+ +
centred on p p = (0.693 x 22.7 1.065 x 40)/(22.7 40) = 0.930 with
Bayesian Methods 275

+
variance 4/(22.7 40) = 0.064 (standard deviation 0.252). For the
enthusiasts, the probability of the hazard ratio being greater than,
say, 1.5 is 1- 4((log(1.5) - 0.930)/0.252) = 1 - 0.019 = 0.981. They
would be happy for the trial to be stopped early.
However, combing these data with the prior of the sceptics gives a
posterior distribution, with mean p p = (0 x 22.7+ 1.065 x 40)/(22.7+
40) = 0.679 and variance 4/(22.7 + 40) = 0.064. This posterior
distribution assigns probability 1 - 4((0.405 - 0.679)/0.252) = 1 -
0.139 = 0.861 to the hazard ratio, being greater than 1.5. This
might not be sufficient to persuade the sceptic that the trial should
be stopped.

9.9. SAMPLE SIZE ESTIMATION FOR


EARLY PHASE TRIALS

The previous section has described Bayesian inferential procedures


that could be used in the design and analysis of a phase I11 trial.
However, in the early, more exploratory stages of the development of
a new therapy, the case for a Bayesian decision theory approach is
stronger. The application of decision theory requires the specification
of a gain function. In some circumstances, the gain function may be
concerned with maximising the relevant information gathered from
a particular experiment. In others it may express on a common scale
the various costs of an experiment and the possible patient suffering
or benefit from the experiment. We briefly consider each of these.
Whitehead and Brunier (1995) consider the context of a phase-I
dose-finding experiment. Patients are treated one-at-a-time and the
response of each patient is observed before the next is treated. They
consider a model in which the probability that the ith patient given
a dose di suffers an adverse reaction follows a logit model of the form:

(9.14)
276 Design and Analysis of Clinical Trials

and a generalisation of the simple conjugate beta distribution (see


Section 9.2) for the prior distribution for 0 is chosen. The objective
of the experiment is specified as finding the dose D* at which an
acceptable probability p of adverse reaction is obtained. For example,
for p = 0.2, then:

10g(0.2/0.8) = log 81 + 02 log D' (9.15)

In order to minimise the variance in the estimate of D*, the gain


function is specified so as to obtain G(8,d) = Var(fi(d)*)-'. The
design action Ai that will maximise the expected gain, given the
information available from the first i patients, is given by:

/ ~ ( 8 7d>h(olvi>do (9.16)

where h(8)yi) is the posterior density of 0 given the observed re-


sponses up to subject i . In general, numerical methods are required
to solve for Ai, the optimal next dose.
Although the choice of dose is data dependent, tending to fall
if previous responses are adverse or rise if not , maximum-likelihood
estimation of dose-response models to the resulting data remains
unnaffected.
Now consider a phase-I1 trial. The decision to undertake a phase
I1 trial should take into account possible subsequent gain and loss.
Brunier and Whitehead (1994) consider the possible gains laid out
in Table 9.3.
In deciding to continue to the phase 111 trial, they argue that it
is intuitively reasonable to continue if a critical number c of the n
phase I1 patients are successfully treated. They then propose a gain
function parametrised in terms of n and c:
Bayesian Methods 277

Table 9.3. Gain Function for a Phase I1 Trial.


Gained Successes.
During During After
Phase I1 Phase I11 Phase I11
n ( P -Po> m ( P -Po) W ( P -Po>

p true probability of success of new treatment.


=
po probability of success on standard treatment.
=
n = number of patients given new treatment in
phase I1 trial.
m = number of patients given each treatment in a
subsequent phase I11 trial.
w = number of patients receiving the recommended
treatment after the phase I11 trial.

where

ho(p) = prior for p

A n ( p , c) = is the probability that the new treatment progresses


from phase I1 t o phase I11

B ( p ) = the power function of the phase I11 trial

Other quantities are defined in Table 9.3.


Summarising the hypothetical results of the phase I11 trial as suc-
cesses S and failures F and total marginal totals T , with subscripts
N and C for new and control treatments, then B ( p ) is given by:

(9.18)
278 Design and Analysis of Clinical Trials

where

and k is the critical normal deviate at the chosen significance level


a.
The analysis may be extended to also consider the costs of trials.
As a consequence of varying the design parameters, they conclude
that the optimal size of the phase I1 trial:

0 decreases with the precision of the prior ho(p)


0 increases with the size of the ‘patient horizon’ w
0 decreases by using a realistic Phase I11 power function B ( p )
0 decreases with increasing phase I1 trial cost
0 increases (generally) with increasing phase I11 trial cost.

The use of a decision analysis approach within the context of


an equivalence trial is discussed by Lindley (1998), contrasting the
conclusions with those reached using standard frequentist methods.

9.10. REPORTING OF RESULTS


Not surprisingly, the Bayesian perspective suggests that the report-
ing of trial results could, with advantage, take a rather different
form to the traditional frequentist one. In conformity with the gen-
eral trend, Parmar et al. (1994) argue against the focus on p-values.
They argue that studies should persuade reasonable sceptics and en-
thusiasts alike. This is particularly important for sequential trials
that are now very common. The use of the sceptical prior acts to re-
duce the risk of premature stopping of a trial in which some benefit is
shown early in the trial. The use of the enthusiastic prior is to reduce
the risk of premature stopping due to early lack of benefit. In both
cases, the amount of ‘pull-back’ is much greater in the early stages
Bayesian Methods 279

of a trial when there is relatively little data as compared to the in-


formation content of the prior. Freedman, Spiegelhalter and Parmar
(1994) suggest that results sections should include the usual tables,
curves? estimates and standard errors based on the data, and that
interpretation sections should describe prior distributions and give
treatment difference estimates (point and interval) using unadjusted
(uninformative prior), sceptical and enthusiasts prior. For each prior,
a table with probabilities of treatment difference falling below, within
and above the range of equivalence should be presented.

9.11. SUMMARY
Although still unfamiliar to some statisticians and regulators, the
use of Bayesian methods in the design and analysis of clinical trials
is becoming increasingly common. We have done little more than
sketch some of the key ideas and procedures in Bayesian estimation
and analysis. With faster computation and improvements in the-
ory and associated algorithms, virtually all statisticians involved in
trials are likely to find themselves using Bayesian technology. To do
so often does not require an acceptance of the use of subjective prob-
ability. However, progress is being made in formalising procedures
that do use subjective probability. When combined with explicit gain
functions into a formal decision framework, the approach is clearly
well suited to guiding the research strat,egy of a group of investgators
sharing a common interest. However, we are yet to see whether these
will gain longer term acceptance in the more public arena of definitive
trials. Bayesian methods are now also much used for meta-analysis,
the subject of the final chapter.
CHAPTER10

Meta- Analysis

10.1. INTRODUCTION

In the Cambridge Dictionary of Statistics in the Medical Sciences,


meta-analysis is defined thus:

A collection of techniques whereby the results of two or more inde-


pendent studies are statistically combined to yield an overall answer to a
question of interest. The rationale behind this approach is to provide a test
with more power than is provided by the separate studies themselves. The
procedure has become increasingly popular in the last decade or so, but
is not without its critics, particularly because of the difficulties of know-
ing which studies should be included and to which population final results
actually apply.

In essence, meta-analysis is a more systematic approach to com-


bining evidence from multiple research projects than the classical
review article. Chalmers and Lau (1993) make the point that both
approaches can be biased, but that at least the writer of a meta-
analytic paper is required by the rudimentary standards of the dis-
cipline to give the data on which the conclusions are based, and
to defend the development of these conclusions by giving evidence
that all available data are included, or to give the reasons for not
including the data. In contrast, the typical reviewer arrives at con-
clusions that may be biased and then selects data to back them up.

280
Meta-Analysis 281

Chalmers and Lau conclude, “It seems obvious that a discipline which
requires that all available data be revealed and included in an analysis
has an advantage over one that has traditionally not presented anal-
yses of all the data on which conclusions are based.”
So meta-analysis is likely t o have an objectivity that is inevitably
lacking in literature reviews and can also achieve greater precision
and generalisability of findings than any single study. Consequently,
it is not surprising that the technique has become one of the great-
est growth areas in medical research. Some examples of its use are
given in Chalmers (1987); Louis: Fienberg and Mosteller (1985), and
DerSimonian and Laird (1986). There remain, however, sceptics who
feel that the conclusions drawn from meta-analysis often go beyond
what the techniques and the data justify. Some of their concerns
are echoed in the last part of the definition with which this chapter
began, and in the following quotation from Oakes (1993);

The term meta-analysis refers to the quantitative combination of data


from independent trials. Where the result of such combination is a de-
scriptive summary of the weight of the available evidence, the exercise is
of undoubted value. Attempts t o apply inferential methods, however, are
subject to considerable methodological and logical difficulties. The selec-
tion and quality of the trials included, population bias and the specification
of the population to which inference may properly be made are problems
to which no satisfactory solutions have been proposed.

Despite such concerns, there has been a striking increase in pub-


lished meta.-analyses of clinical trials. Most stem from the fact that
so many trials are too small for adequate conclusions to be drawn
about potentially small advantages of particular therapies. Advo-
cacy of large trials is a natural response to this situation] but it is
not always possible to launch very large trials before therapies be-
come widely accepted or rejected prematurely. In fact, there are
now several instances of very large trials being started after meta-
analysis of multiple small ones were strongly positive (see Antman
et al., 1992). In this way the use of meta-analysis implies that the
282 Design and Analysis of Clinical %als

whole question of power and sample size determination considered


in Chapter 2 needs to be re-evaluated. In Freiman et al. (1978),
investigators about to undertake a randomised controlled trial are
recommended to consult a biostatistician for help in estimating the
number of patients to obtain a useful answer, and if the answer
were more patients than might be available, to abandon the trial.
In a later paper, however, one of the authors confesses to a dramatic
change of heart and suggests that the second recommendation is one
which is very poor. (See Chalmers and Lau, 1993.)
In this chapter, we shall give a number of examples of meta-
analysis and address some of the issues of most concern when a p
plying the method. We begin with perhaps the most important
component of any proposed meta-analysis, namely the selection of
which studies to include.

10.2. SELECTION OF STUDIES - WHAT


TO INCLUDE
The selection of the studies to be integrated by a meta-analysis will
clearly have a bearing on the conclusions reached. Selection is a
matter of inclusion and exclusion and the judgements required are
often difficult: should only randomised trials be included; should
poor quality research be excluded (and who should judge quality?);
should only a single endpoint be analysed? - and so on. A great
deal of attention has been paid to those issues and there is no con-
sensus view. Chalmers and Lau (1993), aware of the many op-
portunities to change the results of a meta-analysis by selecting
trials the conclusions of which agree with preconceived notions and
reject those that do not, suggest the rather involved process outlined
in Fig. 10.1 for minimising bias. Blinding papers by blotting out the
sources and disguising the results allows ‘quality’ to be scored in an
unbalanced fashion.
Pocock (1993) suggests that the process of selection has three
components: breadth, quality and representativeness. Breadth re-
lates to the decision as to whether to study a very specific narrow
Meta- Analysis 283

EiTltles and UI
References
from completed

Non-RCT L
rejects
Rejects

Copy 1 of full paper ’ Copy 2 of lull paper


and blinded paper
in separate folders
Differentially - and blinded paper
in separate folders
to Reader 1 to Reader 2

Master file
Quality scored with full paper
using Form 6 for ~ ~ ~ ~ ~ ~ ~ ~ r ~ d f o
blinded methods, blinded methods,
Form C tor blinded Form C for blinded
results with full results with full
paper. and Form A paper, and Form A
for blinded data for blmded data
extraction extraction

Rejected it Data Data Rejected 11


found not extracted extracted found not
to meet to meet
screening I screening
requirements 1 Readers have a consensus Conference requirements
2 84 Design and Analysis of Clinical Trials

question (e.g., the same drug, disease and setting for studies following
a common protocol) or a more generic problem (e.g.: a broad class of
treatments for a range of conditions in a variety of settings). Pocock
suggests that the broader the meta-analysis, the more difficulty there
is in interpreting the combined evidence as regards future policy;
consequently, the broader the meta-analysis the more it needs to be
interpreted qualitatively rather than quantitatively,
The reliability of a meta-analysis will depend on the quality of
the data in the included studies. In meta-analyses of clinical tri-
als, for example, adherence to recognised criteria of acceptability
(blinding, randomisation, analysis by intention to treat, etc.) should
be looked for, and any relaxation of such standards requires careful
consideration as to whether the consequent precision of more data
is counter-productive, given the increased potential for loss of cred-
ibility. Determining quality would be helped if the results from so
many trials were not so poorly reported. In the future, this may
be improved by the Consolidation of Standards for Reporting Trials
(CONSORT) statement ( B e g et al., 1996). The core contribution of
the CONSORT statement consists of a flow diagram (see Fig. 10.2)
and a checklist (see Table 10.1). The flow diagram enables review-
ers and readers to quickly grasp how many eligible participants were
randomly assigned to each arm of the t,rial. Such information is fre-
quently difficult or impossible t o ascert,ain from trial reports as they
are currently presented. The checklist identifies 21 items that should
be incorporated in the title, abstract, int.roduction, methods, results
or conclusion of every randomised clinical trial.
Ensuring that a meta-analysis is truly representative can be
problematic. It has long been known that journal articles are not
a representative sample of work addressed to any particular area of
research (see, for example, Sterlin, 1959, Greenwald, 1975 and Smith,
1980). Significant research findings, in particular, are more likely to
find their way into journals than non-significant results. An infor-
mal method of assessing the effect of publication bias is the so-called
funnel plot, in which effect size from a study is plotted against the
Met a- Analysis 285

I Registeredor eligible patients (n = ...I 1


Not randomized (n = ...I

intervention as allocated In = ...I as allwxated In = .. I

Did not receive intervention


as allocated I n = ...)

1
Followed up In = ...I Followed up (n = ...I
Timing of primary and Timing of primary and
secondary outcomes secondary outcomes

Withdrawn (n = ...I
Intervention ineffective In = ...I
Withdrawn (n = .. J
Intervention ineffective In = ...I I
Lost to iollow-up In = ...I Lost to follow-up (n = ...)
Other In = ...I Other I n = ...I
I
Completed trial (n = ...I

Fig. 10.2. Flow diagram of CONSORT statement.

Table 10.1. Checklist for the CONSORT Statement.

Head. Subhead. Descriptor Was it On what


reported? page number?
Title Identify the study as a
randomised trial.
Abstract Use a structured format.
Meth. Protocol Describe
Planned study population,
together with inclusion/
exclusion criteria.
Planned interventions and
their timing.
Primary and secondary
outcome measure(s) and
the minimum important
difference(s) and indicate
286 Design and Analysis of Clinical Trials

Table 10.1. (Continued)

Head. Subhead. Descriptor Was it On what


reported? page number?
indicate how the target
sample size was projected.
Rationale and methods
for statistical analyses,
detailing main comparative
analyses and whether they
were completed on an
intent ion-to- t reat basis.
Prospectively defined
stopping rules (if
warranted).
Assignment Describe
unit of randomisation
(e.g., individual, cluster,
geographic).
Method used t o generate
the allocation schedule.
Method of allocation,
concealment and timing
of assignment.
Method to separate the
generator from the executor
of assigment .
Masking Describe mechanism (e.g.,
(Blinding) capsules, tablets); similarity
of treatment characteristics
(e.g., appearance, taste)
allocation schedule control
(location of code during
Meta- Analysis 287

Table 10.1 ( Continued )

Head. Subhead. Descriptor Was it On what


reported? page number?
trial and when broken);
and evidence for successful
blinding among participants,
person doing intervention
outcome assessors and
data analysts.
Results: Provide a trial profile
(figure)
Participant summarising participant
flow and flow, numbers and timing
follow-up of randomisation assign-
ment, interventions and
measurements for each
randomised group.
Analysis State estimated effect of
intervention on primary
and secondary outcome
measures, including a
point estimate and
measure of precision
(confidence interval).
State results in absolute
numbers where feasible
(e.g., 10/20, not 50%).
Present summary data and
appropriate descriptive and
inferential statistics in
sufficient detail to allow
for alternative analyses
and replication.
288 Design and Analysis of Clinical Trials

Table 10.1 (Continued)

Head. Subhead. Descriptor Was it On what


reported? page number?
Describe prognostic
variables by treatment
group and any attempt
t p adjust for them.
Describe protocol deviations
form the study as planned,
together with the reasons.
Comment State specific interpretation
of study findings, including
sources of bias and
imprecision (internal
validity) and discussion of
external validity, including
appropriate quantitative
measures when possible.
State general interpretation
of the data in light of the
totality of the available
evidence.

study’s sample size. Because of the nature of sampling variability,


this plot should, in the absence of publication bias, have the shape
of a pyramid with a tapering ‘funnel-like’ peak. Publication bias
will tend to skew the pyramid by selectively excluding studies with
small, non-significant effects. Such studies predominate when the
sample sizes are small, but are likely to be increasingly less common
as the sample sizes increase. Consequently, their absence removes
part of the left hand corner of the pyramid. The effect is illustrated
in Fig. 10.3.
Meta- Analysis 289

160 - Y

120 -
X

wIn
.-
X

80-
5
0
0

40 - x o o 0 0
X 0

:0 0 0
0
0 x i O?dQ
0,"= O O
XP X X % ; F
0 I " I . . ' 1 . 1 I ' * I " I

160 - X

120 -
X
a,
N
.-
v) X
$ 80-
E
UY
X

40 - X
X X
X
X
xmxx
x
xx x xfX * x : r
0 , - . [ I T - , , . 1 , 7 . I ) C ' V l

(b)
Fig. 10.3. Funnel plot.
290 Design and Analysis of Clinical Trials

A meta-analysis that relies solely upon published data is liable


to be biased, and confidence in the overall quality of a meta-analysis
will be greatly enhanced if explicit efforts are made t o identify all
studies and not just those that are published. This can often be a
formidable enterprise possibly taking several years, but there is little
doubt that such efforts considerably reduce the danger of overesti-
mating treatment efficacy.

10.3. THE STATISTICS OF


META- AN ALYSIS

A meta-analysis should almost always begin with some form of graph-


ical display. A single figure displaying point estimates and confidence
intervals for the individual studies is often sufficient to show, at a
glance, what the broad conclusions of the analysis are. Figure 10.4,
for example, shows the results from a series of randomised controlled
trials of endoscopic treatment of bleeding peptic ulcer (see Sacks
et al., 1990). The binary outcome variable in each case was whether
or not there was a reduction in recurrent or continued bleeding. The
results are presented as confidence intervals for the odds ratios. Al-
though four of the 25 trials favour the control group, the overall effect
of endoscopic therapy appears to be highly statistically significant.
For some meta-analyses, a diagram such as Fig. 10.4 might be
so clear that more detailed statistical investigation becomes unnec-
essary. In some circumstances, however, investigators might be in-
terested in a global test of significance for the overall null hypothesis
of no effect in all studies, or, more commonly, an overall estimate
of the magnitude of the effect. Concentrating on the latter, sup-
pose that there are N studies t o be analysed. There are two dis-
tinct approaches to providing an estimate of overall effect size in a
meta-analysis. The first (and most commonly used) is to take the
N studies as the only ones of interest. The second is to regard
the N studies as a sample from a larger population of relevant stud-
ies. The two approaches correspond, respectively, t o the assumptions
Meta-A nalysis 291

Odds Ratio 95% CI


Sludy
1 Vallon
-
Year

1980
0.01 0.I 1 10
- 100

2 Swain 1981
3 Pam 1982
4 Rurgeens 1982
5 MacLeOd 1983
6 Jensen 1984
7 Kernohan 1984
8 Goudie 1984
9 Freilas 1985
10 Swain 1986
1 1 0Brten 1986
12 Krels 1987
13 Brearley 1987
14 Moreto 1987
15 Laine 1987
16 Panes 1987
17 Chung 1987
18 Balanzo 1988
19 Fellenon 1989
20 Angerinas 1989
21 Rutgeens 1989
22 Chozzini 1989
23 Lame 1989
24 Acabvxhi 1990
25 Oxner 1992

Overall
Favours Treatment Favours Control

Fig. 10.4. Confidence intervals for odds ratios from a series of randomised
trials of endoscopic treatment of bleeding peptic ulcer (from Sacks et al.,
1990).

of a fixed and random set of studies for the analysis of variance.


Table 10.2 presents the basis of estimation for each approach for a
general measure of effect size.
The fixed effect estimate is reliant on the strong assumption
that there is no true heterogeneity between studies, i.e., they are all
estimating the same true effect and only differ because of sampling
variation. Such an assumption can be tested (see Table l0.2), but the
292 Design and Analgsis of Clinical Trials

Table 10.2. The Statistical Basis of Meta-Analysis.

(a) Fixed enects

0 Let Y denote the generic measure of t,he effect of an experimental inter-


vention (the eflect size}.
0 Let W denote the reciprocal of the variance of effect size.
0 Under the assumption of a fixed set of studies, a n estimator of the as-
sumed common underlying effect size is:
N

i= 1

The standard error of this estimator is:

[irl I
-1/2

SE.(Y) =

0 An approximate l O O ( 1 - a ) % confidence interval for the population effect


size, say $, is:

I
N I
N

(b) Random effects

0 Under the assumption that the studies are a random sample from a larger
population of studies, there is a mean population effect size, say 4, about
which the study-specific effect sizes vary. Thus, even if each study’s
results were based on sample sizes so large that the standard errors of the
Y s were zero, there would still be study-to-st,udy variation because each
study would have its own underlying effect size (i.e., its own parameter
value).
Met a-Analysis 293

Table 10.2. (Continued)

0 Let D denote the variance of the studies' effect sizes (a quantity yet to
be determined) and let Q denote the following statistic for measuring
study-to-study variation in effect size:

N
Q = CW,(y,
-Y ) 2

i= 1

a The estimated component of variance due t o interstudy variation in effect


size, D , is calculated as:

D=O ifQ<N-1

D = [Q- ( N - l)]/U if Q >N - 1


(see Der Simonian and Laird, 1986) where

u = ( N - 1) [w - 34
with W and S& being the mean and variance of the Ws.
0 An approximate l O O ( 1 - a)% confidence interval for I$ is:

I
N I
N

where

i= 1
294 Design and Analysis of Clinical %als

Table 10.2. (Continued)


The confidence interval for the underlying parameter is wider in the ran-
dom effects model than in the fixed effects model. A random effects
analysis suggest more uncertainty in estimating the underlying parame-
ter than does a fixed effects analysis.
Test for homogeneity of studies - the test statistic; Q is given by:

N
Q = CWz(y,- Y ) Z
a= 1

The hypothesis of a common effect size is rejected if Q exceeds at


the chosen significance level.

test lacks power. Consequently, when such a test is non-significant


we cannot assert that the fixed effect method is correct but only
that major heterogeneity is not present. But in most applications of
meta-analysis, homogeneity is a priori implausible given the variety
of study designs.
The random effects method attempts to incorporate statistical
heterogeneity into the overall estimate of an average effect. Accord-
ing to Fleiss (1993) “the random effects model anticipates better than
the fixed effects model the possibility that some studies not included
in the analysis are underway, are about to be published, or perhaps
have even already been published in a obscure journal, and that the
results in some of the non-included studies are different from the re-
sults in most of the meta-analysed studies”. Again in a report on
combining information sponsored by the National Research Council’s
Committee on Applied and Theoretical Statistics (National Research
Council, 1992) the conclusion is that “modeling would be improved
by an increase in the use of random effects models in preference to
the current default of fixed effects models.”
But Pocock (1993) suggests that the random effects method has
both conceptual and practical problems; he dislikes both the as-
sumptions that the studies are a random sample from a hypothetical
Meta-Analysis 295

population of such studies, and that if heterogeneiby exists, that


substantially more weight is given to the smaller studies than in the
fixed effect approach when smaller studies are often those of poorer
quality. The problem is usually not too serious provided that the
number of patients in a given trial is not too small - the problem
becomes less not as the number of trials increases but as the number
of patients per trial increases.
Thompson (1993) also raises some doubts about the random
effects model, arguing that it may mask the true reasons for the un-
derlying heterogeniety and that possible explanations should be in-
vestigated. One possibility would be to use generalised linear models
(see Chapters 4 and 7) allowing for both fixed study level covariates,
indicating characteristics of the study that. may be related t o effect
size and thought to explain some of the heterogeneity, together with
a random component to accomodate unexplained heterogeneity.
DeMets (1987) and Bailey (1987) discuss the strengths and weak-
nesses of the two competing models. Bailey, for example, suggests
that when the research question involves extrapolation into the future
- Will the treatment have an effect, on the average? - then the

random effects model for the studies is the appropriate one. This
research question implicitly assumes that there is a population of
studies from which those analysed in the meta-analysis were sam-
pled, and anticipates future studies being conducted or previously
unknown studies being uncovered. In contrast, when the research
questions concern whether treatment has produced an effect, on the
average in the set of studies being analysed, then the fixed effects
model is the one-tc-use. Meier (1987) and Pet0 (1987) present fur-
ther arguments in favour of one or other of the models.
Oakes (1993) also considers the methodological arguments that
can arise in the application of meta-analysis, but in addition, ad-
dresses the wider question of whether the statistical combination of
results from different trials is legitimate at all. He makes the point
that the principal feature of statistical inference is the argument from
sample to population. For the argument to have pretensions to strict
296 Design and Analysis of Clinical Trials

legitimacy, the sample must be the result of a random process; for


the argument to make sense, the population must be identifiable. In
the case of meta-analysis, Oakes argues that there are grounds for
concern on both counts, and at the heart of the debate as to whether
and to what extent data from differing studies may legitimately be
combined - the so-called problem of apples and oranges - is the
question of “what is the population?” Without a clear conception of
the population, Oakes suggests that a parameter or its meta-analytic
estimate is meaningless, and devoid of application. Oakes finally
summarises his own views about applying the inferential process in
meta-analysis as follows:

It is my guess that the controversial inferential aspect of meta-analysis,


like the positivist philosophy to which it is the witting or unwitting heir,
will not fall to sceptical critiques such as mine but rather it will collapse
under its own weight in the conscientious application of its practitioners.
Researchers will move effortlessly from the insightful collection, stratifi-
cation and summary of disparate results t o the intelligent interpretation of
them. The specious pseudo-objective accuracy of the intervening statistical
inferences will be discreetly dropped.
Medical science will not suffer.

Table 10.3. Results from Nine Trials Comparing SMFP and NaF
Toothpastes in Prevention of Caries Development.

Study NaF SMFP SMFP-NaF


n Mean SD n Mean SD
~~ ~~ ~ ~ ~~

1 134 5.96 4.24 113 4.72 4.72 +0.86


2 175 4.74 4.64 151 5.07 5.38 +0.33
3 137 2.04 2.59 140 2.51 3.22 +0.47
4 184 2.70 2.32 179 3.20 2.46 +0.50
5 174 6.09 4.86 169 5.81 5.14 -0.28
6 754 4.72 5.33 736 4.76 5.29 0.04
7 209 101.0 8.10 209 10.90 7.90 +OX0
8 1151 2.82 3.05 1122 3.01 3.32 +0.019
9 679 3.88 4.85 673 4.37 5.37 +0.49
Meta-Analysis 297

Like Thompson a.nd Pocock (1991), it appears tha.t Oakes views


meta-analysis as largely an objective descriptive technique and plays
down the value of the quantitative results of such an analysis. But
the growing use of meta-analysis in a number of speciality areas with
quantitative results being regarded at least as importantly as those
that are qualitative, argues that in practice meta-analysis is seen as
more substantial than a mere descriptive technique.

10.4. TWO EXAMPLES OF THE


APPLICATION OF META-ANALYSIS
10.4.1. The Comparison of Sodium
Monofluorophosphate (SMFP) and
Sodium Fluoride (NaF) Toothpastes in
the Prevention of Caries Development
Table 10.3 shows the results of nine randomised trials comparing
SMFP and N a F toot,hpastes for the prevention of caries development.
The outcome in each trial was the change, from baseline, in the
decayed, missing (due to caries) and filled surface (DMFS) dental
index. Calculation of the test for heterogeneity given in Table 10.2
leads to a value for Q of 5.40. Testing this as a chi-squared with
8 degrees of freedom provides no evidence of heterogeniety in the
nine trials; consequently, only the fixed effects model will be used.
Applying the general methodology described in Table 10.2 to the
mean differences leads to a pooled estimate of the effect size of 0.283
with a 95% confidence interval of (0.103,0.464). The trials provide
convincing evidence of a greater change from baseline in the DMFS
index when using toothpastes containing ShlFP than when using
pastes containing NaF.

10.4.2. The Effectiveness of Aspirin in


Preventing Death after a Myocardial
Infarct ion
Table 10.4 shows the results of seven randomised trials, in chronologi-
cal order, of the effectiveness of aspirin (versus placebo) in preventing
298 Design and Analysis of Clinical Trials

Table 10.4. Results of Seven Trials of Aspirin for Preventing


Death after a Myocardial Infarction.
Survived Died Total
1. Elwood et al. Aspirin 566 49 615
(1974) Placebo 557 67 624
Total 1123 116 1239
2. Coronary drug Aspirin 714 44 758
project group (1976) Placebo 707 64 77
Total 1421 108 1529
3. Elwoodand Aspirin 730 102 832
Sweetnam (1979) Placebo 724 126 850
Total 1454 228 1682
4. Breddin et al. Aspirin 285 32 317
(1979) Placebo 271 38 309
Total 556 70 626
5. Persantine-Aspirin Aspirin 725 85 810
Study Group (1980) Placebo 354 52 406
Total 1079 137 1216
6. Aspirin study Aspirin 2021 346 2267
Group (1980) Placebo 2038 219 2257
Total 4059 465 4524
7. ISIS-2 Collaborative Aspirin 7017 1570 8587
Group (1988) Placebo 6880 1720 8600
Total 13897 3290 17187

death after a myocardial infarction (the data are taken from Fleiss,
1993). In this case, a suitable measure of effect size is the logarithm of
the odds ratio (see Chapter 4). The meta-analysis results are given
in Table 10.5. The fixed effects model indicates a positive effect
for aspirin, but using the random effects approach, the confidence
interval for the overall odds ratio expands to include the value one,
indicating no effect.
Meta- Analysis 299

Table 10.5. A Meta-analysis of the Data in Table 10.4.


(a) Fixed effects
Study Odds Ratio log(odds ratio) W = l/VAR[log(or)]
1 1.18 0.17 24.09
1 1.47 0.38 24.29
3 1.25 0.22 48.80
4 1.25 0.22 15.44
5 1.25 0.23 27.41
6 0.88 -0.12 103.99
7 1.12 0.11 663.92

Using log(or) as the measure of effect size we have:

= 0.104

and a 95% confidence interval for the logarithm of the population odds ratio
is (0.039,0.169).
With respect to the odds ratio itself, the point estimate is 1.109 and the
confidence interval is (1.040,1.184).

(b) Random effects


Again using log(or) as the measure of effect size, the various quantities
defined in Table 10.1 are as follows:

W = 129.85

S& = 56363.38

U = 345.02

D = 0.008

Q = 8.76

Y' = 0.1158

W; ... W,* = 20.19,20.33,35.09,17.74,23.14,56.74,105.11


300 Design and Analysis of Clanical Trials

The 95% confidence for the logarithm of the population odds ratio is
(-0.0025,0.2341).
With respect to the odds ratio itself, the point estimate is 1.123 and the
confidence interval is (0.997,1.263).

10.5. BAYESIAN META-ANALYSIS

A Bayesian approach to meta-analysis might begin with the following


random effects model:

(10.1)
82 - N(6,T2) 2 = 1,.. . , N (10.2)

where 13i is the true but unknown effect in the ith study and 4 is
the unknown population effect; it is the latter quantity which is of
most interest, since it represents the pooled effect indicated by the
studies. Finally, g2 is the population variance: or the between-study
variability, and is also of interest since it is a measure of how variable
the effect is within a population. In the case where T~ = 0, a fixed
effect model is obtained.
In a Bayesian setting, prior distributions would be needed for
the unknown parameters of the model. A normal prior is typically
specified for 417. Although a prior for r could be specified, for ex-
ample specifying the precision, l / ~as, being gamma distributed (see
Section 9.3 for such a specification for the variance of a random effect
for frailty), such a specification is not necessary. With a prior for 7,
n(7>,the posterior distribution for T given the observed effect sizes
f(71Y) is proportional to:

+
where Si = r2 (o?/ni)
Assuming a uniform distribution for T ( T ) , the posterior dist.ribu-
tion can then be approximated by a discrete distribution calculated
Meta-Analysis 301

over, say, 100 equally spaced points that straddle its mode and with
a range sufficient to include all the values in which the density is a t
least 1%of that at the mode (Hedges, 1998). The simple sum of the
100-point densities provides the normalising constant.
With f ( ~ 1 Yapproximated
) by this discrete distribution, a. pos-
terior density for 4 J Y is easily calculated as a finite mixture of the
conditional densities g(qIY,T ) . This conditional density is Normal
with mean:
(10.4)

and variance
a 2 ( r )= c-kj’ (10.5)

giving the posterior density for 4 as:


(10.6)

where it(.) is the standard normal probability density function.


Monte Carlo methods provide an alternative means of calculating
g ( $ y ) . A value of T is first sampled from ~ ( T J Yand
) , then a value
for $ is drawn from f($lY, T).
A plot of the estimated posterior density g(4IY)provides a useful
graphical expression for the uncertainty in 4. A mean and variance
can be calculated as, too, can the probability that the mean effect
size is positive, or above some other clinically meaningful value.
In principle, the Bayesian approach has the advantage over the
methods of the previous section for random effects models, of more
completely accounting for the uncertainty in the estimation of
the random effect in the standard errors and confidence intervals
of the overall effect estimate. In practice, the results calculated
using Bayesian methods differ little from those using the methods
of the previous sections unless the number of studies is very small
or the studies are very variable. It is also under these circumstances
302 Design and Analysis of Clinical Trials

that the particular choice of prior for 7 may have a substantial im-
pact. It is therefore important to examine the sensitivity of findings
to different choices of prior.
Kass and Wasserman (1996) provide some useful background for
the choice of prior distributions. For T = 0, the Bayesian estimates
of the study specific effects are all equal to the common fixed effect
estimate. When T is very large in proportion to the within study sam-
pling variance, then the Bayesian estimates approach their observed
values. For intermediate values of T ,the estimates lie between their
observed values and the overall common mean - they are ‘shrunken’
towards the mean, the studies with the more uncertain effect sizes
(usually the smaller studies) being more shrunken. Therefore pri-
ors for T that are more concentrated on zero will generate greater
shrinkage than more diffuse priors.
However, Bayesian methods are relatively simply modified to con-
sider more general random effects specifications, e.g., to examine the
implications of using study effect size distributions with heavier tails,
such as Student’s t or gamma distributions (e.g., Smith et al., 1995).
Analyses with an informative prior for the effect size are also a little
more complex.

10.6. META-ANALYSIS IN MORE


COMPLEX SETTINGS

The analyses above have assumed that the contributing studies and
their corresponding effect sizes Y , are exchangeable. No study-level
covariates have been considered and are conditional on the 13i and
the Y , being independent. Non-independence may arise through a
clustering of contributing trials by centre, by an overlap of patients
across trials, or sometimes by the presence of single studies that
provide more than one estimate of effect. In the more complex cases,
the question arises as to whether a more effective analysis might be
achieved by combining data at the level of the individual subject,
i.e., data pooling, rather than at the level of the trial. Senn (1998)
Met a-Analysis 303

Odds Ralio 95% C!


01 02 0.5 I 2 5 10

1
1
1
1
1
1
1
1
'I
I
1
1
1
1

L i
Favours Treatment Favours Control

Fig. 10.5. Confidence intervals for odds ratios from a series of randomised
trials of endoscopic treatment. of bleeding peptic ulcer presented in cumu-
lative form (form Chalmers and Lau, 1993).

considers this question in relation to covariate adjustment, favouring


data pooling.

10.7. SUMMARY

Meta-analysis has become one of the greatest growth areas in medical


research, despite the many tough criticisms that have been levelled
304 Design and Analysis of Clinical Trials

at the approach. (An excellent recent account is available is Nor-


mand, 1999.) Users (and potential users) of meta-analysis need to
keep in mind the problems that arise at each stage - what studies
to select for inclusion; should fixed or random effects models be used
to provide an overall estimate of effect size; to what population can
results be generalised? And if meta-analysis, despite its wide usage,
remains controversial, what can be said of the suggestion of Chalmers
and Lau (1993) that cumulative meta-analysis might be employed to
establish such strong evidence of a treatment effect to make further
randomised trials unjustified? Figure 10.5, for example, shows the
data previously presented in Fig. 10.4 in a cumulative fashion. Note
that by 1982 the chances that endoscopic therapy was not reducing
recurrent or continued bleeding, when compared with standard ther-
apy, was less than one in a thousand. Yet in the 25th published trial,
patients were randomly assigned to a control group between April
1989 and June 1991. Chalmers and Lau comment, “If the authors of
the later publications knew of the previous ones, how much proof did
they require, or were they tied to the null hypothesis yoke?” They
conclude, “the advent of cumulative meta-analysis makes one worry
about the present custom of conducting clinical trials as if no prior
trials of the same treatment had been done. When is it no longer
ethical to assign patients at random to a control group in a new
definitive large trial?”
APPENDIX

Software

Many of the methods described in this book could not be applied


routinely without the aid of some convenient piece of statistical soft-
ware. Fortunately, many excellent statistical packages are now widely
available, although our intention in this appendix is not to attempt
either a comprehensive review or critique. Instead, we shall give
details of packages that we feel are most suitable for specific applica-
tions and analyses. Links to web pages for a wide variety of statistical
software can be found at:
http:/ /www .st ata.com/support /links/stat software.html

1. SOFTWARE FOR SAMPLE SIZE


DETERMINATION

nQuery Advisor Version 2.0: A package for choosing an appropriate


sample size for many types of design and analysis including two-
sample and paired t-tests, analysis of variance, crossover designs,
nonparametric tests, survival analysis and regression. Very sim-
ple to use and well described is the user manual written by Janet
D. Elashoff.
Available from:

0 Statistical Solutions Ltd. 8 South Bank, Crosse’s Green, Cork,


Ireland: Tel +353 21 319629, Fax +353 21 319630:
http://www.statsol.ie:

305
306 Design and Analysis of Clinical Dials

0 In the USA, Boston, MA, USA 1-800-262-1171:


ht tp://www .statsolusa.com: info@statsolusa.com

MS-DOS 3.1 or higher, Microsoft Windows 3.1, Windows 95 or


Windows N T is required.
A large number of other software packages for estimating sample
size when designing studies can be found on:
http://www.interchg.ubc.ca/cacb/power

2. SOFTWARE FOR MULTIPLE


IMPUTATION OF MISSING VALUES

Software is available from Statistical Solutions (address above) and


from:
http://www.stat .psu.edu/-jls/misoftwa.html

3. SOFTWARE FOR INTERIM ANALYSIS

0 East, Version 2 : Cytel Software Corporation, 675 Massachu-


setts Ave., Cambridge, MA 98195, USA.
0 PEST3, Planning and Evaluation of Sequential Trials: MPS
Research Unit, University of Reading, Earley Gate, Reading
RG6 6FN, United Kingdom. Both these packages are reviewed
in Emerson (1996).

4. SOFTWARE FOR GENERALISED


LINEAR MODELS

Generalised linear models can be fitted using most statistical pack-


ages, including:

0 SAS - SAS Institute Inc., SAS Campus Drive, Cary, NC


27513, USA.
Appendix 307

0 STATA - Stata Corp., 702, University Drive East, College


Station, TX 77840, USA.
0 GENSTAT - Numerical Algorithms Group Ltd., Wilkinson
House, Jordon Hill Rd., Oxford OX28DR, UK.
0 SPSS - SPSS Inc., 444 N. Michigan Ave., Chicago, IL 60611,
USA.

5. SOFTWARE FOR ANALYSING


LONGITUDINAL DATA
For the summary measure or response feature approach described in
Chapter 5, any of the standard packages can be used. For the more
involved modelling approaches covered in Chapters 6 and 7 there are
a number of possibilities:

5.1 Multivariate normal regression model


Both BMDP5V and SAS PROC MIXED (see, Getting Started with
PROC M I X E D , SAS Institute Inc.) can be used to fit the types
of models described in Chapter 6. A variety of structures may be
specified for the covariance matrix of the repeated measurements,
including independence, compound symmetry, random effects and
unstructured.

4.1 Generalized Linear Mixed Models (GLMM)


Three general approaches to GLMM estimation are available.
(1) Maximum likelihood estimation:
0 Tractable special cases; binomial mixed logistic in EGRET
(Statistics and Epidemiology Research Corp., 1107 NE 45th
St., Suite 520, Seattle, WA 98105); negative binomial of
Stata’s nbreg.
0 Standard numerical integration; gllamm written for
STATA. (Computational time can be considerable.)
308 Design and Analysis of Clinical D-ials

(2) An approximation based on a linearising adaptation to stan-


dard random effects algorithms.
0 This is available in SAS (glimmix) - downloadable from:
‘‘http://www.sas.com/techsup/download/samples/
stat -and-iml”
with documentation by Wolfinger, Reference for glimmix: a
tutorial on mixed models (SAS Institute Inc.).
0 Also available in MLW in - the Multilevel Models Project,
Mathematical Sciences, Institute of Education, University
of London, 20 Bedford Way, London WClHOAL, UK.
(3) A Bayesian approach using Monte Carlo Markov Chain
estimation. Models may be specified and fitted entirely
within an MCMC framework (e.g., BUGS available from
http://www.mrcbsu.cam.ac.uk) or MCMC may be used
as a final stage (e.g., MLWin).

5.2 Generalised Estimating Equations approach


This approach is now available in SAS, STATA and S-PLUS -
Statsci Division, MathSoft Inc, 1700 Westlake Avenue N, #500 Seat-
tle, Washington 98109-9891, USA.

5.3 Missing Data

(1) The Diggle/Kenward model for dropouts is implemented in


the OSWALD (Object oriented software for the analysis of
longitudinal data in S) written by David M. Smith, Bill
Robertson and Peter Diggle. The software runs under S or
S-Plus on both Unix and DOS systems. OSWALD is avail-
able from:
http://www.maths.lancs.ac.uk:2080/~maa036/oswald/
Appendix 309

(2) Weighting approaches for the treatment of missing data can be


implemented in SAS, SUDAAN (Research Triangle Institute,
P.O. Box 12194, Research Triangle Park, NC 27709, USA),
and particularly simply in STATA.

6. SOFTWARE FOR BAYESIAN ANALYSIS


BUGS provides a model specification language allowing a wide vari-
ety of models to be estimated via Gibbs sampling. The documenta-
tion provides an extensive range of examples.

7. SOFTWARE FOR SURVIVAL ANALYSIS


(1) The routine methods of analysis, such as Kaplan-Meier, Cox's
regression, etc., are available in all standard packages such as
SAS and SPSS.
(2) A wider range of techniques are implemented in S-Plus and
STATA.
(3) A number of other very useful methods are described in
Extending the Cox Model, Technical Report Number 58, Mayo
Clinic, Rochester, Minnesota, written by Professor Terry
Therneau.
REFERENCES

Adkinson, N.F., Eggleston, P.A., Eney, D., Goldstein, E.O., Schuberth,


K.C., Bacon, J.R., Hamilton, R.G., Weiss, M.E., Arshand, H., Mein-
ert. C.L., Tonascra, J. and Wheeler, B. (1997). A controlled trial of
immunotherapy for asthma in allergic children, New Engl. J. Med.
336: 324-331.
Altman, D.G. and De Stavola, B.L. (1994). Practical problems in fitting
a proportiona.1 hazards model to data with updated measurements of
the covariates, Statist. Med. 13: 301-341.
Amberson, J.B., McMahon, B.T. and Pinner, M. (1931). A clinical trial
of sanocrysin in pulmonary tuberculosis, Am. Rev. Tubercul. 24:
401-435.
Andersen, B. (1990). Methodological Errors in Medical Research, Blackwell
Scientific Publications, Oxford.
Anderson, P.K., Borgan, O., Gill, R.D. and Keiding, N. (1993). Statistical
Models Based on Counting Processes, Springer-Verlag, NY.
Anderson, P.K. and Gill, R.D. (1982). Cox’s regression model for counting
process: a large sample study, Ann. Statist. 10: 1100-1120.
Anscombe, F.J. (1954). Fixed-sample size analysis of sequential observa-
tions, Biometrzcs 10: 89-100.
Antman, EM., Lau, J., Jimenez-Silva, J., Kupelnick, B., Mosteller, F.
and Chalmers, T.C. (1992). A comparison of results of meta-
analysis of randomized control trials and recommendat,ions of clinical

311
312 Design and Analysis of Clinical Trials

experts; treatments for myocardial infarction, J. Am. Med. Assoc.


268: 240-248.
Armitage, P., McPherson, C.K. and Rowe, B.C. (1969). Repeated signifi-
cance tests on accumulating data, J. Roy. Statist. SOC.Ser. A 132:
235-244.
Armitage, P. (1983). Trials and errors: the emergence of clinical statistics,
J. Roy. Statist. SOC.Ser. A 146: 321-334.
Aspirin Myocardia Infarction Study Research Group (1980). A randomised
controlled trial of aspirin in persons recovered from myocardial infarc-
tion. J . Am. Med. Assoc. 243: 661-669.
Azzalini, A. and Cox, D.R. (1984). Two new tests associated with analysis
of variance, J. Roy. Statist. SOC.Ser. B., pp. 335-343.
Bailey, K.R. (1987). Inter-study differences: How should they influence the
interpretation and analysis of results? Statist. in Med. 6: 351-358.
Barlow, W.E. and Prentice, R.L. (1988). Residuals for relative risk regres-
sions. Biometrika 75: 65-74.
Beecham, J. and Knapp, M.R.J. (1992). Costing psychiatric interventions.
In Measuring Mental Health Needs (eds. G. Thornicroft, C. Brewin
and J. Wing), Gaskell, London, pp. 163-183.
Begg, C., Cho, M., Eastwood, S., Morton, R., Moher, D., Ohlein, I., et al.
(1996). Improving the quality of reporting of randomised controlled
trials; the CONSORT statement, J. Am. Med. Assoc. 276: 637-939.
Berkey, C.S., Anderson, J.J. and Hoaglen, D.C. (1996). Multiple-outcome
meta analysis of clinical trials, Statist. Med. 15: 537-557.
Berry, D.A. (1990). Basic principles in designing and analyzing clinical
studies. In Statistical Methodology in the Pharmaceutical Sciences (ed.
D.A. Berry), Marcel Dekker, New York, pp. 1-55.
Berry, D.A. (1993). A case for Bayesianism in clinical trials, Statist. Med.
12: 1377-1393.
Berzuni, C. (1996). Medical monitoring. In Markov Chain Monte Carlo
in Practice (eds. W.R. Gilks, S. Richardson and D.J. Spiegelhalter),
Chapman and Hall, London, pp. 321-338.
Binder, D.A. (1983). On the variances of asymptotically normal estimators
from complex surveys, Int. Statist. Rev. 51: 279-292.
Blair, R.C., Troendle, J.F. and Beck, R.W. (1996). Control of familywise
errors in multiple endpoint assessments via stepwise permutation tests,
Statist. Med. 15: 1107-1121.
References 313

Bracken, M.B. (1987). Clinical trials and the acceptance of uncertainty,


Br. Med. J. 294: 111-112.
Bradford Hill, A. (1962). Statistical Methods in Clinical and Preventative
Medicine, Livingstone, Edinburgh.
Breddin, K., Loew, D., Lechner, K. and Uberla, E.W. (1979). Secondary
prevention of myocardial infarction. Comparison of acetylsalicykic
acid, phenprcoumon and placebo. A multicenter, two-year prospective
study. Thrombosis and Haemostasis 41: 225-236.
Breslow, N.E. (1974). Covariance analysis of censored survival data, Bio-
metrics 30: 89-99.
Breslow, N.E. (1990). Tests of hypotheses in overdispersion regression and
other quasi-likelihood models, J. Am. Statist. Assoc. 8 5 : 565-571.
Bruner, H.C. and Whitehead, J. (1994). Sample sizes for phase I11 clinical
trials derived from Bayesian Decision Theory, Statist. Med., 2493-2502.
Byar, D.P. (1985). Assessing apparent treatment - covariate interactions
in randomized clinical trials, Statist. Med. 4: 255-263.
Cain, K.C. and Lange, N.T. (1984) Approximate case influence for the
proportional hazards regression model with censored data, Biometrics
40: 493-499.
Carey, V., Zeger, S.L. and Diggle, P. (1993). Modelling multivariate binary
data with alternating logistic regressions, Biometrika 80: 517-526.
Chalmers, T.C. (1987). Meta-analysis in clinical medicine, Trans. Am.
Clin. Climatol. Assoc. 99: 144-150.
Chalmers, T.C. and Lau, J. (1993). Meta-analysis stimulus for change in
clinical trials, Statist. Meth. Med. Res. 2: 161-172.
Chalmers, T.C., Matta, R.J., Smith, H. and Kunzler, A.M. (1977). Evi-
dence favouring the use of anticoagulents in the hospital phase of acute
myocardial infarction, N. Engl. J. Med. 297: 1091-1096.
Chastang, C., Byar, D. and Piantadosi, S. (1988). A quantitative study
of the bias in estimating the treatment effect caused by omitting a
balanced covariate in survival models, Statist. Med. 7: 1243-1255.
Chen, C.-H. and Wang, P.C. (1991). Diagnostic plots in Cox’s regression
model, Biometrics 47: 841-850.
Chesher, A. and Jewitt, I. (1987). The bias of a heteroskedasticity consis-
tent covariance matrix estimator, Econometrica 5 5 : 1217-1222.
Chesher, A and Lancaster, T . (1983) The estimation of models of labour
behaviour, Rev. &con. Stud. 50: 609-624.
314 Design and Analysis of Clinical Trials

Chuang-Stein, C. (1993). The regression fallacy, Drug Info. J. 27: 1213-


1220.
Chuang-Stein, C. and Tong, D.M. (1997). The impact and implication of
regression to the mean on the design and analysis of medical investi-
gations, Statist. Meth. Med. Res. 6: 115-128.
Clayton, D.G. (1974). Some odds ratio statistics for the analysis of ordered
categorical data, Biometrika 61: 525-531.
Cnaan, A., Laird, N.M. and Shasor, P. (1997). Using the general linear
mixed model to analyse unbalanced repeated measures and longitudi-
nal data, Statist. Med. 16: 2349-2380.
Collett, D. (1991). Modelling Binary Data, Chapman and Hall, London.
Collett, D. (1994). Modelling Survival Data in Medical Research, Chapman
and Hall, London.
Collins, R. (1993). Capopril, nitrates and magnesium after myocardial
infraction: Inetroduction and results - ISIS-4. Quoted from taped
presentation. American Heart Association’s 66th Scientific sessions,
Atlanta, 8-11 November.
Coronary Drug Project Research Group. (1973). The Coronary Drug Pro-
ject: Design, methods and baseline results. 47 (supplement I), 1-1-1-50.
Coronary Drug Project Research Group. (1976). Asprin in coronary heart
disease, J. Chronic Diseases, 29: 625-642.
Cornfield, J . (1976). Recent methodological contributions to clinical trials,
A m . J. Epidemiol. 104: 408-421.
Cox, D.R. (1972). Regression models and life tables, J. Roy. Statist. SOC.
Ser. B34: 187-220.
Cox, D.R. (1975). Partial likelihood, Biometrika 62: 269-276.
Cox, D.R. (1998). Discussion, Statist. Med. 17: 387-389.
Cox, D.R. and Oakes, D. (1984). Analysis of Survival Data, Chapman and
Hall, London.
Cox, D.R. and Snell, E.J. (1968). A general definition of residuals. J. Roy.
Statist. SOC.Ser. B30: 248-275.
Cramer, J.A., Collins, J.F. and Mallson, R.H. (1988). Can categorization of
patient background problems be used to determine early termination
in a clinical trial? Contr. Clin. Trials 9: 47-63.
Crombie, A.C., (1952). Avicenna’s influence on the medieval scientific tra-
dition. In Avicenna: Scientist and Philosopher (ed. G.M. Wickens),
Luzac and Co. Ltd., London, pp. 84-107.
References 315

Crouchley, R. and Pickles, A. (1993). Information matrix tests of speci-


fication for univariate and multivariate proportional hazards models,
Biometrics 49: 1067-1076.
Crowley, J. and Hu, M. (1977). Covariance analysis of heart transplant
survival data. J. Am. Statist. Assoc. 78: 27-36.
Crowley, J . and Storer, B.G. (1983). Comment on a re-analysis of the
‘Stanford Heart Transplant Data’ by Aitkin, Laird and Francis, J.
Am. Statist. Assoc. 78: 277-281.
Curran, W.J. and Shapiro, E.D. (1970). Law, Medicine and Forensic Sca-
ence, 2nd ed., Little, Brown and Co, Boston.
Curran, D., Bacchi, M., Schmitz, S.F.H., Molenberghs, G. and Sylvester
(1998). Identifying the types of missingness in quality of life cancer
clinical trials, Statist. Med. 17: 739-756.
Davis, C.S. (1991). Semi-parametric and non-parametric methods for t.he
analysis of repeated measurements with applications t o clinical trials,
Statist. Med. 10: 1959-1980.
De Mets, D.L. (1987). Methods for combining randomized clinical trials:
Strengths and limitations, Statist. Med. 6: 341-348.
De Mets, D.L. and Lan, K.K.G. (1994). Interim analysis - the alpha
spending function approach, Statist. Med. 13: 1341-1352.
Der Simonian, R. and Laird, N. (1986). Meta-analysis in clinical trials,
Contr. Glin. Trials 7:177-186.
Diggle, P.J. (1992). Discussion of Liang. K.-Y., Zeger, S.L. and Qaqish,
B. (1992). Multivariate regression analyses for categorical data (with
discussion), J. Roy. Statist. SOC.B54: 3-40.
Diggle, P.J. (1988). An approach to the analysis of repeated measures,
Biometrics 44: 959-971.
Diggle, P.J. (1998). Dealing with missing values in longitudinal studies.
In Statistical Analysis of Medical Data: New Developments (eds. B.S.
Everitt and G. Dunn), Arnold, London, pp. 203-228.
Diggle, P.J. and Kenward, M.G. (1994). Informative drop-out in longitu-
dinal analysis (with discussion), AppE. Statist. 43: 49-93.
Diggle, P.J., Liang, K.Y. and Zeger, S.L. (1994). Analysis of Longitudinal
Data, Oxford Scientific Publications, Oxford.
Donner: A. (1984). Approaches to sample size estimation in the designing
of clinical trials - a review, Statist. Med. 3: 199-214.
316 Design and Analysis of Clinical Trials

Donner, A. and Klar, N. (1994). Cluster randomization trials in epidemi-


ology - theory and application, J. Statist. Plan. Infer. 42: 37-56.
Drum, M. and McCullagh, P. (1993). Comment on Fitzmaurice, G.M.,
Laird, N. and Rotnitzky, A. Regression models for discrete longitudinal
responses, Statist. Sci. 8 : 284-309.
DuMouchel, W.H. (1990). Bayesian meta-analysis. In Statistical Method-
ology in Pharmaceutical Sciences (ed. D.A. Berry), Marcel Dekker,
New York, pp. 509-29.
Dunn, G. (1989). Design and Analysis of Reliability Studies: Statistical
Evaluation of Measurement Errors, Arnold. London.
Dunn, G. (1999). Compliance in clinical trials, Statist. Med., in press.
Dunn, G., Everitt, B.S. and Pickles, A. (1993). Modelling Cowariances and
Latent Variables using EQS, Chapman and Hall, London.
Efron, B. (1979). Bootstrap methods: another look at the jackknife, Ann.
Statist. 7: 1-26.
Efron, B. (1998). Foreword in special issue on Analyzing Non-Compliance
in Clinical Trials, Statist. Medi. 17: 249-250
Efron, B. and Feldman, D. (1991). Compliance as a n explanatory variable
in clinical trials, J. Am. Statist. Assoc. 86: 9-17.
Efron, B. and Tibshirani, R. (1986). Bootstrap measures for standard
errors, confidence intervals and other measures of statistical accuracy,
Statist. Sci. 1: 54-77.
Elwood, P.C., Cochrane, A.L., Burr, M.L., Sweetnam, P.M., Williams,
G., Welsby, E., Hugher, S.J. and Renton, R. (1974). A randomized
controlled trial of acetyl solicyclic acid in its secondary prevention of
mortality from myocardial infarction, Brit. Med. J., 1: 436-440.
Elwood, P.C. and Sweetnam, P. (1979). Aspirin and secondary mortality
after myocardial infarction, Lancet 2: 1313-1315.
Emerson, S.S. (1996). Statistical packages for group sequential methods,
The A m . Statist. 50: 183-192.
Everitt, B.S. (1996). An introduction to finite mixture distributions,
Statist. Meth. Med. Res. 5 : 107-127.
References 317

Fairclough, D.L., Peterson, H.F. and Chang, V. (1998). Why are miss-
ing quality of life data a problem in clinical trials of cancer therapy?
Statist. Med. 17: 667-678.
Falissard, B. and Lellouch, J . (1992). A new procedure for group sequential
analysis in clinical trials, Biornetrics 48: 373-388.
Farewell, V.T.(1982). The use of mixture models for the analysis of survival
data with longterm survivors. Biometrics 38: 1041-46.
Farewell, V., Fleming, T.R. and Harrington, D.P. (1991). Counting Pro-
cesses and Survival Analysis, John Wiley and Sons, New York.
Feinstein, A.R. (1991). Intention-to-treat policy for analyzing randomized
trials: Statisticad distortions and neglected clinical challenges, Chap-
ter 28. In Patient Compliance in Medical Practice and Clinical Trials
(eds. by J.A. Cramer and B. Spilker), Raven Press Ltd, New York.
Finkelstein, D.M. and Schoenfeld, D.A. (1994). Analysing survival in the
presence of an auxiliary variable, Statist. Med. 13: 1747-1754.
Finney, D.J. (1990). Repeated Measurements: what is measured and what
repeats? Statist. Med. 9: 639-644.
Firth, D. (1987). On the efficiency of quasi-likelihood estimation, Bio-
metrika 74: 233-246.
Fisher, R.A. and MacKenzie, W.A. (1923). Studies in crop variation: 11.
The manurial response of different potato varieties, J. Agri. Sci. 13:
311-320.
Fitzmaurice, G.M., Laird, N. and Rotnltzky, A. (1993). Regression models
for discrete longitudinal responses, Statist. Sci. 8: 2844309.
Fleiss, J.L. (1986). The Design and Analysis of Clinical Experiments,
Wiley, New York.
Fleiss, J.L. (1993). The statistical basis of meta-analysis, Statist. Meth.
Med. Res. 2: 121-145.
Follman, D. (1995). Multivariate tests for multiple endpoints in clinical
trials. Statist. Med. 14: 1163-1175.
Fox, W., Sutherland, I. and Daniels, M. (1954). A five-year assessment of
patients in a controlled trial of streptomycin in pulmonary tubercolo-
sis, Quart. J . Med. 23: 347.
Freedman, L.S. (1982). Tables of the number of patients required in clinical
trials using the logrank test, Statist. Med. 1: 121-130.
318 Design and Analysis of Clinical Trials

Freedman, L.S. and Spiegehalter, D.J. (1983). The assessment of subjective


opinion and its use in relation t o stopping rules for clinical trials. The
Statist. 32: 153-161.
Freedman, L.S., Spiegehalter, D.J. and Parmar, M.K.B. (1994). The what,
why and how of Bayesian Clinical trials monitoring, Statist. Med. 13:
1371-1384.
Freidman, L.M., Furberg, C.D. and De Mets, D.L. (1985). Fundamentals
of Clinical Trials, 2nd ed., PSB Publishing, Littleton, MA.
Freiman, J.A., Chalmers, T.C. and Smith, H. (1978). The importance of
beta, the type I1 error and sample size in the design and interpretation
of the randomized control trial: survey of ‘negative’ trials, N. Engl. J.
Med. 299: 690-694.
Freirich, E., Gehan et al. (1963). The effect of 6-mercaptopurine on the
duration of steroi-induced remissions in acute leukemia; a model for
evaluation of other potentially useful therapies, Blood 21: 699-716.
Frison, L. and Pocock, S.J. (1992). Repeated measures in clinical trials:
analysis using mean summary statistics and its implications for design,
Statist. Med. 11: 1685-1704.
Gail, M.H. and Simon, R. (1985). Testing for qualitative interactions be-
tween treatment effects and patient subsets, Biornetrics 41: 361-372.
Gardner, M.J. and Altman, D.G. (1986). Confidence intervals rather than
p-values: estimation rather than hypothesis testing, Br. J. Med. 292:
746.
Gardner, M.J. and Altman, D.G. (1989). Statistics with Confidence, British
Medical Journal, London.
Gehan, E.A. (1984). The evaluation of therapies: Historical control studies,
Statist. Med. 3: 315-324.
Gelman, A. (1996). Inference and monitoring convergence. In Marlcov
Chain Monte Carlo in Practice (eds. W.R. Gilks, S. Richardson and
D.J. Spiegelhalter), Chapman and Hall, London, pp. 131-144.
Giardiello, F.M., Hamilton, S.R., Krush, A.J., Piantadosi, S., Hylind, L.M.,
Celano, P., Booker, S.V., Robinson, C.R. and Offerhaus, G.J.A.
(1993). Treatment of colonic and rectal adenomas with sulindac in
familial adenomatous polyposis, N. Engl. J. Med. 328: 1313-1316.
Gilks W.R., Richardson, S. and Spiegelhalter, D.J. (1996). Introducing
Markov Chain Monte Carlo. In Marlcov Chain Monte Carlo in Practice
References 319

(eds. W.R. Gilks, S. Richardson and D.J. Spiegelhalter), Chapman


and Hall, London, pp. 1-20.
Gilks, W.R., Clayton, D.G., Spiegelhalter, D.J., Best, N.G., n/lcNeil,
A.J., Sharples, L.D. and Kirby, A.J. (1993). Modelling complexity:
applications of Gibbs sampling in medicine, J. Roy. Statist. SOC.
B55:39-52.
Gill, R. and Schumacher, M. (1987). A simple test of the proportional
hazards assumption, Biometrika 74: 289-300.
Glonek, G.F.V. and McCullagh, R. (1995). Multivariate logistic models,
J . Roy. Statist. SOC.Ser. B57: 533-546.
Godambe, V.P. (1960). An optimal property of regular maximum likelihood
estimation, Ann. Math. Statist. 31: 1209-1211.
Goetghebeur, E.J.T. and Shapiro, S.H. (1996). Analysjng non-compliance
in clinical trials: Ethical imperative or mission impossible? Statist.
Med. 15: 2813-2826.
Goldstein, H. (1991). Nonlinear multilevel models, with an application to
discrete response data, BiometP-ika 78:45-51.
Gornbein, J.A., Lazaro, C.G. and Little, R.J.A. (1992). Incomplete data
in repeated measures analysis, Statist. Meth. Med. Res. I: 275-295.
Gould, A.L. (1998). Multi-centre trial analysis revisited, Statist. Med. 17:
1779-1 797.
Greenwald, A.G. (1975). Consequences of prejudice against the null hy-
pothesis, Psychol. Bull. 85: 845-857.
Grossman, J., Paramar, M.K.B., Spiegelhalter, D.J. and Freedman, L.S.
(1994). A unified method for monitoring and analysing controlled
trials, Statist. Med. 13: 1815-1826.
Hastle, T.J. and Tibshirani, R.J. (1990). Generalized Additive Models,
Chapman and Hall, London and New York.
Heart Special Project Committee. (1988). Organization, review and ad-
ministration of cooperative studies (Greehberg Report}: A report. from
the Heart Special Project Committee t o the National Advisory Coun-
cil, May 1967, Contr. Clin. Trials 9: 137-148.
Hedges, L.V. (1998). Bayesian meta-analysis. In Statistical Analysis of
Medical Data: New Developments (eds. B. Everitt and G. Dunn),
Arnold, New York, pp. 251-276.
320 Design and Analysis of Clinical Trials

Heyting, A., Tolboom, J.T.B.M. and Essers, J.G.A. (1992). Statistical


handling of drop-outs in longitudinal clinical trials, Statist. Med. 11:
2043-2061.
Hjalmarson, A., Elmfeldt, D., Herlitz, J., Holmberg, S., Nyberg, G., Ryden,
L., Swedberg, K., Waagstein, F., Waldenstram, A., Vedin, A., Wedel,
H., Wilhelmsen, L. and Wilhelmsson, C. (1981). A double blind trial
of metoprolol in acute myocardial infarction - effects on mortality,
Circulation 64: 140.
Hochberg, T. (1988). A sharper Bonferroni procedure for multiple tests of
significance, Biometrika 75: 800-803.
Holm, S. (1979). A simple sequentially rejective multiple test procedure.
Scand. J. Statist. 6: 65-70.
Holtzman, J.L., Kiam, D.C., Berry, D.C., Mottonen, L., Barrett, G., Harri-
son, L.I. and Conrard, G.J. (1987). The pharmacodynamic and phar-
macokinetic interaction of flecouride acetate and propranolol; effects
on cardiac function and drug clearance, Eur. J. Clan. Pharmacol. 33:
97-99.
Hsieh, F.Y., Crowley, J. and Tormey, D.C. (1983). Some test statistics for
use in multistate survival analysis, Biometrika 70: 111-119.
Huber, P. (1967). The behaviour of maximum likelihood estimators under
nonstandard conditions, Proc. Fifth Berkeley Symp. Math. Statist.
Probab. 1, 221-233. University of California Press, Berkeley.
ISIS-2 Collaborative Group (1988). Randomized trial of intravenous strep-
tokinase, oral aspirin, both or neither among 17 187 cases of suspected
acute myocardial infarction. ISIS-2. Lancet 2: 349-360.
Jennison, C. and Turnbull, B. (1989). Interim analysis: The repeated con-
fidence interval approach (with discussion), J. Roy. Statist. Soc. Ser.
B51: 305-362.
Jones, B. and Kenward, M.G. (1989). Design and Analysis of Cross-over
Trials, Chapman and Hall, London.
Jones, B., Teather, D., Wang, J. and Lewis, J.A. (1998). A comparison of
various estimators of a treatment difference for a multi-centre clinical
t,rial, Statist. Med. 17: 1767-1777.
Kalbfleisch, E.L. and Prentice, R.L. (1973). Marginal likelihood based on
Cox’s regression and life model, Biometrika, 60: 267-278.
Kalbfleisch, J.D. and Prentice, R.L. (1980). The Statistical AnaZysis of
Failure Time Data, Wiley, New York.
References 321

Kaplan, E.L. and Meier, P. (1958). Nonparametric estimation from incom-


plete observation, J. A m . Statist. Assoc. 53: 457-481.
Karim, M.R. (1989). Technical Report #674, Dept. of Biostatistics, John
Hopkins University.
Kass, R.E. and Wasserman, L. (1996). The selection of prior distributions
by formal rules, J. Am. Statist. Assoc. 91: 1434-1470.
Kenward, M.G. and Jones, B. (1992). Alternative approaches to the anal-
ysis of binary and categorical repeated measurements, J. Blophamal.
Studies 2: 137-170.
Kenward, M.G., Lesaffre, E. and Molenberghs, G. (1994). An application
of ML and GEE to the analysis of ordinal data from a long study with
cases. Biornetrics, 50: 945-953.
Kenward, M.G. (1987). A method for comparing profiles of repeated mea-
surements, Appl. Statist. 36: 296-308.
Knatterud, G.L., Rockhold, F.W., George, S.L., Barton, F.B., Davis, C.E.,
Fairweather, W.R., Honohan, T., Mowery, R. and O'Neill, R. (1998).
Guidelines for quality assurance in multicenter trials: A position pa-
per, Cont. Clan. Trials 19: 477-93.
Kuipers, E., Fowler, D., Garety, P., Chisholm, D., Freeman, D., Dunn,
G., Bebbington, P. and Hadley, C. (1998). London-East Anglia ran-
domised controlled trial of cognitive-behavioural therapy for psychosis
I11 Follow-up and economic evaluation at 18 months, Br. J. Psychiat.
173: 61-68.
Lagokos, S.W. (1977). Using auxiliary variables for improved estimates of
survival time, Biornetrics 33: 399-404.
Laird, N.M. (1988). Missing data in longitudinal studies, Statist. Med. 7:
305-31 5.
Lan, K.K.G. and De Mets, D.L. (1983). Discrete sequential boundaries for
clinical trials, Biornetrika 70:659-663.
Lapaucis, A., Sackett, D.L. and Roberts, R. (1988). An assessment of
clinically useful measures of the consequences of treatment, N . Engl.
J. Med. 26: 1728-1733.
Lavori, P.W. (1992). Clinical trials in psychiatry: Should protocol deviation
censor patient data? Neuropsychophamnacology 6: 39-48.
Lee, E.T. (1992). Statistical Methods for Survival Data Analysis, Wiley,
New York.
322 Design and Analysis of Clinical Trials

Lee, Y.J. (1983). Quick and simple approximations of sample sizes for
comparing two independent binomial distributions: Different-sample
size case, Biometrics 40: 239-242.
Lee, Y.J., Ellenberg, J.H., Hirtz, D.G. and Nelson, K.B. (1991). Analysis
of clinical trials by treatment actually received: is it really and option?
Statist. Med. 10: 1595-1605.
Levine, R.J. (1981). Ethics and Regulations of Clinical Research, Urban
and Schwartzenberg, Baltimore.
Liang, K.-Y. and Zeger. S.L. (1986). Longitudinal data analysis using
generalized linear models, Biometrika, 73: 13-22.
Liang, K.-Y., Zeger, S.L. and Qaqish, B. (1992). Multivariate regression
analyses for categorical data (with discussion), J. Roy. Statist. SOC.
B54: 3-40.
Lin, D. and Wei, L.J. (1989). The robust inference for the Cox proportional
hazards model, J. Am. Statist. Assoc. 84: 1074-1078.
Lind, J. (1753). A Treatise of the Scurvey reprinted in Lind’s Treatise on
Scvrwey (eds. C.P. Stewart and D. Guthrie), Edinburgh University
Press, Edinburgh, (1953). Sands, Murray, Cochran, Edinburgh.
Lindley, D.V. (1998). Decision analysis and bio-equivalence trials, Statist.
Sci. 13: 136-141.
Lindsey, J.K. (1993). Models f o r Repeated Measurements, Oxford Science
Publications, Oxford.
Lindsey, J.K. and Lambert, P. (1998). On the appropriateness of marginal
models for repeated measurements in clinical trials, Statist. Med. 28:
447-469.
Lipsitz, S.R. and Fitzmaurice, G.M. (1996). Estimating equations for mea-
sures of association between repeated binary responses, Biometrics 52:
903-91 2.
Lipsitz, S.R., Laird, M. and Harrington, D.P. (1991). Generalized estimat-
ing equations for correlated binary data: using the odds ratio as a
measure of association, Biometrika 78: 153-160.
Lipsitz, S.R., Fitzmaurice, G.M., Orav, E.J. and Laird, N.M. (1994). Per-
formance of generalized estimating equations in practical situations,
Biometrics 50: 270-278.
Little, R.J.A. (1995). Modelling the dropout mechanism in repeated mea-
sures studies, J. Am. Statist. Assoc. 90: 1112-1121.
References 323

Little, R.J.A. and Rubin, D.B. (1987). Statistical Analysis with Missing
Data, Wiley, New York.
Louis, T.A., Fienberg, H.V. and Mosteffer, F. (1985). Findings for public
health for meta-analysis, Ann. Rev. Public Health 6: 1-20.
Lubsen, J. and Pocock, S.J. (1994). Factorial trials in cardiology: Pros and
cons, Eur. Heart J. 15: 585-588.
Machin, D. (1994). Discussions of The What, Why and How of Bayesian
Clinical Trials Monitoring, Statist. Med. 13: 1385-1390.
Malmberg, K. (1997). Prospective randomized study of intensive insulin
treatment on long term survival after acute myocardial infarction in
patients with diabetes mellitus Br. Med. J. 314: 1512-1515.
Mantel, N. and Haenszel, W. (1959). Statistical aspects of the analysis of
dates from retrospective studies of disease, J. Natl. Cancer Inst. 22:
719-748.
Marubini, E. and Valsecchi, M.G. (1995). Analysing Survival Data from
Clinical Trials and Observational Studies, Wiley, New York.
Matthews, J.N.S. (1993). A refinement to the analysis of serial data using
summary measures, Statist. Med. 12: 27-37.
Matthews, J.N.S. (1994). Discussion of Diggle and Kenward, Appl. Statist.
43: 49-93.
Matthews, J.N.S., Altman, D.G., Campbell, M.J. and Royston, P. (1989).
Analysis of serial measurements in medical research, Br. Med. J. 300:
23&235.
Mauritsen, R.H. (1984). Logistic regression with random effects, Unpub-
lished PhD thesis, Department of Biostatistics, University of Wash-
ington.
McCullagh, P. (1983). Quasilikelihood functions, Ann. Statist. 11: 59-67.
McCullagh, P. and Nelder, J.A. (1989). Generalized Linear Models, 2nd
ed., Chapman and Hall, London.
McHugh, R.B. and Lee, C.T. (1984). Confidence estimation and the size
of a clinical trial, Contr. Clan. Trials 5 : 157-164.
McPherson, K. (1982). On choosing the number of interim analyses in
clinical trials, Statist. Med. 1: 25-36.
Medical Research Council. (1931). Clinical trials of new remedies (anno-
tations), Lancet 2: 304.
Meinert, C.L. (1986). Clinical Trials - Design, Conduct and Analysis,
Oxford University Press, New York.
324 Design and Analysis of Clinical Trials

Meir, P. (1987). Commentary, Statist. Med. 6: 329-331.


Moertel, C.G., Fleming, T.R., MacDonalds and others. (1990). Levamisole
and Fluorairacil for adjustment therapy of resected colon carcinoma,
N . Engl. J. Med. 332: 353-358.
Moreau, T., O’Quigley, J. and Lellouch, J. (1986). On D. Schoenfeld’s
approach for testing the proportional hazards assumption, Biometrika
73: 513-515.
Muthen, B.O. (1984). A general structural equation model with dichoto-
mous, ordered categorical and continuous latent variable indicators,
Psychol. Med. 49: 115-132.
National Institutes of Health: NIH Almanac. Publication number 81-5,
Division of Public Information, Bethesda, MC.
Neale, M.C. (1997). Mx Statistical Modeling. Box 126 MCV, Richmond,
VA 23298: Department of Psychiatry. 4th Edition.
Neten, A. and Dennett, J. (1996). Unit Costs of Health and Social Care.
Canterbury: Personal Social Services Research Unit, University of
Kent.
Normand, S.T. (1999). Meta-analysis: Formulating, evaluating, combining
and reporting, Statist. Med. 18: 321-359.
Oakes, M. (1986). Statistical Inference: A Commentary for the Social and
Behavioural Sciences, John Wiley and Sons, Chichester.
Oakes, M. (1993). The logic and role of meta-analysis in clinical research,
Statist. Meth. Medi. Res. 2: 147-160.
O’Brien, P.C. and Fleming, T.R. (1979). A multiple testing procedure for
clinical trials, Biometrics 35: 549-556.
Oldham, P.D. (1962). A note on the analysis of repeated measurements of
the same subjects, J. Chron. Dis. 15: 969-977.
O’Quigley, J. and Pessione, F. (1989). Score tests for homogeneity of regres-
sion effect in the proportional hazards model, Biornetrics 45: 135-144.
Pampailona, S. and Tsiatis, A.A. (1994). Group sequential designs for
one-sided and two-sided hypothesis testing with provision for early
stopping in favour of the null hypothsis, J. Statist. Plan. Infer. 42:
19-35.
Parmar, M.K.B., Spiegelhalter, D.J. and Freedman, L.S. (1994). The
CHART trials: design and monitoring, Statist. Med. 13: 1297-1312.
Patterson, H.D., and Thompson, R. (1971). Recovery of inter-block infor-
mation when block sizes are unequal, Biometrika 58: 545-554.
References 325

Patulin Clinical Trials Committee. (1994). Clinical trial of Patulin in the


common cold, Lancet 2: 373-375.
Peduzzi, P., Wittes, J . and Detre, K. (1993). Analysis-as-randomized and
the problem of nonadherence - an example from the veterans affairs
randomized trial of coronory-artery bypass surgery, Statist. Med. 12:
1185-1195.
Persantine-Aspirin Reinfarction Study Research Group (1980). Persantine
and aspirin in coronary heart disease, Circulation 62: 449-461.
Peterson, B. and Harrell, F.E. (1990). Partial proportional odds models for
ordinal response variables, Appl. Statist. 39: 205-217.
Peto, R. (1982). Statistical aspects of cancer. In Treatment of Cancer (ed.
K.E. Halnan), Chapman and Hall, London, pp. 867-871.
Peto, R. (1987). Why do we need systematic overviews of randomized
trials? Statist. Med. 6: 223-240.
Peto, R., Pike, M.C., Armitage, P. et al. (1976). Design and analysis of ran-
domized clinical trials requiring prolonged observation of each patient:
I Introduction and Design, Br. J. Cancer 34: 585-612.
Piantadosi, S. (1997). Clinical Trials: A Methodological Perspective, Wiley,
New York.
Pickles, A.R. (1987). The problem of initial conditions. In Methods for
Longitudinal Data Analysis (ed. R. Crouchley), Avebury, Aldershot,
pp. 129-149.
Pickles, A.R. (1991). The problem of initial condition in longitudinal anal-
ysis. In Crouchley, R. (ed.) Longitudinal Data Analysis, Aldershot,
Averbury, pp. 129-149.
Pickles, A. and Crouchley, R. (1994). Generalizations and applications of
frailty models for survival and event data, Statist. Meth. Med. Res.
3: 263-278.
Pickles, A. and Crouchley, R. (1995). A comparison of frailty models for
multivariate survival data, Statist. Med. 14: 1447-1461.
Pinheiro, J.C. and De Mets, D.L. (1997). Estimating and reducing bias in
group sequential designs with Gaussian independent increment struc-
ture, Biometrika 84: 831-845.
Pocock, S.J. (1977). Group sequential methods in the design and analysis
of clinical trials, Biometrika 62: 191-199.
Pocock, S.J. (1983). Clinical Trials, John Wiley and Sons, Chichester.
326 Design and Analysis of Clinical Trials

Pocock, S.J. (1992). When to stop a clinical trial, Br. Med. J. 305: 235-
244.
Pocock, S.J. (1996). Clinical Trials: A statistician’s perspective. In Ad-
vances in Biometry (eds. P. Armitage and H.A. David), Wiley, Chich-
ester.
Pocock, S.J. and Abdalla, M. (1998). The hope and the hazards of using
compliance data in randomized controlled trials, Statist. Med. 17:
249-250.
Pocock, S.J., Geller, N.L. and Tsiatis, A.A. (1987). The analysis of multiple
end-points in clinical trials, Biometrics 43: 487-498.
Pregibon, D. (1981). Logistic regression diagnostics, Ann. Statist. 9: 705-
724.
Prentice, R.L. (1973). Exponential survivals with coronary and explanatory
variables, Biometrika 60: 279-288.
Prentice, R.L. (1988). Correlated binary regression with covariates specific
to each binary observation, Biometrics 44: 1033-1048.
Prentice, R.L. (1989). Surrogate endpoints in clinical trials: Definitions and
operations criteria, Statist. Med. 8: 431-440.
Prentice, R.L. and Zhao, L.P. (1991). Estimating equations for parame-
ters in means and covariances of multivariate discrete and continuous
responses, Biometrics 47: 825-883.
Qaqish, B.F. and Liang, K.-Y. (1992). Marginal models for correlated
binary responses with multiple classes and multiple levels of nesting,
Biometrics 48: 939-950.
Rabe-Hesketh, S. and Everitt, B.S. (1999). The effect of misspecifying the
covariance structure when analysing longitudinal data. In preperation.
Reid, N. and Crepeau, H. (1985). Influence functions for proportional haz-
ards regression model. Biometrika 72: 1-9.
Ridout, M.(1991). Testing for random dropouts in repeated measurement
data, Biometrics 47: 1617-1621.
Roberts, G. (1996). Markov chain concepts related to sampling algorithms
Markov Chain Monte Carlo in Practice (eds. W.R., Gilks, S., Richard-
son and D.J., Spiegelhalter), Chapman and Hall, London, pp. 45-58.
Robins, J.M. (1998). Correction for non-compliance in equivalence trials,
Statist. Med. 17: 269-302.
Robins, J.M. and Rotnitsky, A. (1992). Recovery of information and ad-
justment for dependent censoring using surrogate markers. In Aids
References 327

Epidemiology: Methodological Issues (eds. N., Jewell, K., Dietz and


V., Farewell), Birkhauser Press.
Robins, J.M., Rotnitzky, A. and Zhao, L.P. (1995). Analysis of semipara-
metric regression models for repeated outcomes in the presence of
missing data, J. A m . Statist. Assoc. 90: 106-121.
Rogers, W.H. (1993). Regression standard errors in clustered samples, Stata
Technical Bulletin 13: 19-23. Reprinted in Stata Technical Bulletin
Reprints, Vol. 3, pp. 88-94.
Rosner, G. and Tsiatis, A.A. (1989). The impact that group sequential tests
would have made on ECOG clinical trials, Statist. Med. 8: 505-516.
Sacks, H.S., Chalmers, T.C. and Smith, H. (1983). Sensitivity and speci-
ficity of clinical trials: randomized v’s historical controls, Arch. Inter.
Med. 143: 753-775.
Sacks, H.S., Chalmers, T.C., Blum, A.L., Berrier, J . and Pagano, D. (1990).
Endoscopic hemost asis: an effective therapy for bleeding peptic ulcers,
J. A m . Med. Assoc. 262: 494-499.
Sasieni, P. (1993). Maximum weighted partial likelihood estimators for the
Cox Model, J. A m . Statist. Assoc. 88: 135-144.
Schluchter, M.D. (1988). Analysis of incomplete multivariate data using
linear models with structured covariance matrices, Statist. Med. 7:
317-324.
Schoenfield, D.A. (1980). Chi-squared goodness-of-fit tests for the propor-
tional hazards regression model, Biometrika 67: 145-153.
Schoenfield, D.A. (1982). Partial residues for the proportional hazards
regression model, Biometrika 69: 239-241.
Schoenfield, D.A. (1983). Sample size formula for the proportional hazards
regression model, Biometrika 39: 499-503.
Schulz, K.F., Chalmers, F., Grimes, D.A. and Altman, D.G. (1994). As-
sessing the quality of randomization from reports of controlled
trials in obstetrics and gynecology journals, J. A m . Med. Assoc. 272:
125-128.
Schwartz, D. and Lellouch, J . (1967). Explanatory and pragmatic attitudes
in therapeutic trials, J . Chron. Disorders 20: 637-648.
Segal, M.R. and Neuhaus, J.M. (1993). Robust inference for multivariate
survival data. Statist. Med. 93: 1019-31.
Senn, S. (1993). Cross-over Trials in Clinical Research, John Wiley and
Sons, Chichester .
328 Design and Analysis of Clinical Trials

Senn, S. (1994a). Repeated measures in clinical trials: analysis using mean


summary statistics and its implications for design, Statist. Med. 13:
197-198.
Senn, S . (1994b). Testing for baseline balance in clinical trials, Statist.
Med. 13: 1715-1726.
Senn, S. (1997). Statistical Issues in Drug Development, John Wiley and
Sons, Chichester.
Senn, S. (1998). Some controversies in planning and analysing multi-centre
trials, Statist. Med. 17: 1753-1765.
Sieh, F.Y. (1987). A simple method of sample size calculation for inequal-
sample-size deisgns that use the logrank or t-test, Statist. Med. 6:
577-582.
Silverman, W.A. (1985). Human Experimentation: A Guided Step into the
Unknown, Oxford Medical Publications, Oxford.
Simon, R. (1982). Randomization and research strategy, Cancer Treatment
Reports 66: 1083-1087.
Simon, R. (1991). A decade of progress in statistical methodology for clin-
ical trials, Statist. Med. 10: 1789-1817.
Simon, R. (1994). Some practical aspects of the interim monitoring of
clinical trials, Statist. Med. 13: 1401-1409.
Skinner. (1994). Discussion of Diggle and Kenward, Appl. Statist. 43:
49-93.
Smith, D.M. and Diggle, P.J. (1994). OSWALD Version 2: Object Ori-
ented Software for the Analysis of Longitudinal Data in S. Dept. of
Mathematics and Statistics, University of Lancaster, England.
Smith, M.L. (1980). Publication bias and meta-analysis, Eval. Educ. 4:
22-24.
Smith, T.C., Spiegelhalter, D.J. and Thomas, A. (1995). Bayesian ap-
proaches to random-effects metal-analysis: a comparative study,
Statist. Med. 14: 2685-99.
Snell, E.J. (1964). A scaling procedure for ordered categorical data, Bio-
metrics 20: 592-607.
Souhami, R.L. (1994). The clinical importance of early stopping of
randomized trials in cancer treatments, Statist. Med. 13: 1293-1295.
Souhami, R.L. and Whitehead, J. (eds.) (1994). Workshop on Early
Stopping Rules in Cancer Clinical Trials, Statist. Med. 13: 1289-
1500.
References 329

Spiegelhalter D. J. Thomas, A., Best, N. and Gilks, W-. (1996). BUGS 0.5
Examples. Vol. 1. MRC Biostatistics Unit, Cambridge, England.
Spiegelhalter, D.J., Freedman, L.S. and Parmar, M.K.B. (1994). Bayesian
approaches to randomized trials, J. Roy. Statist. SOC.Ser. A157:
357-4 16.
Stanley, K.E. (1980). Prognostic factors for survival in patients with inop-
erable lung cancer. J. Nat. Can.cer Inst. 65: 25-32.
Sterlin, T.D. (1959). Publication decisions and their possible effects in
inferences drawn from tests of significance - or vice-versa, J . Am.
Statist. Assoc. 54: 30-34.
Starer, B.E. and Crowley, J. (1985). A diagnostic for Cox regression and
general conditional likehoods. J. Am. Statist. Assoc. 80: 139-147.
Sutton, H.G. (1865). Ca.ses of rheumatic fever, Guys Hospital Report, 11,
pp. 392428.
Tango, T. (1998). A mixture mode1 to classify individual profiles of re-
peated measurements. In Data Science, Classification and Related
Methods (eds. C . Hayashi, N. Oksumi, K. Yajima, Y. Taraka, H.H.
Bock and Y. Baba), Springer, Tokyo, pp. 247-254.
Thall, P.F. and Vail, S.C. (1990). Some covariance models for longit,udinal
count data with overdispersion, Biometrics 46: 657-671.
Therneau, T.M., Grambsch, P.M. and Fleming, T.R. (1990). Martingale
hazard regression models and the analysis of censored survival data,
Biometrika 77:147-160.
Thompson, S.G. and Pocock, S.J. (1991). Can meta-analysis be trusted?
Lancet 338: 1127-1130.
Thompson, S.G. (1993). Controversies in meta-a.nalysis: the case of the
trials of serum cholesterol reduction, Statist. Meth. hled. Res. 2:
173-192.
Troxel, A.B., Fairclough, D.L., Curran, D. and Hahn, E.A. (1998). Sta-
tistical analysis of quality of life with missing data in cancer clinical
trials, Statist. hfed. 17: 653-666.
Tsiatis A.A. (1981). The asymptotic joint distribution of the efficient
scores test for the proportional hazards model calculated over time,
Biometrika 68: 311-315.
Unit,ed States Congress. (1962). Drug amendments of 1962, Public Law,
87-781, ~ 1 5 2 2Washingt.on.
,
330 Design and Analysis of Clinical Trials

Urquhart, J . and de Klerk, E. (1998). Contending paradigms for the inter-


pretation of data on patient compliance with therapeutic drug regimes,
Statist. Med. 17: 251-268.
Waterhouse, B. (1800). A Prospect of Exterminating the Small Pox, Cam-
bridge University Press, Cambridge.
Waterhouse, B. (1802). A Prospect of Exterminating the Small Pox, Part 11,
Cambridge University Press, Cambridge.
Waterhouse, D.M., Calzone, K.A., Mele, C. and Brenner, D.E. (1993).
Adherence of oral tamoxifen: A comparison of patient self-report, pill
count and microelectronic monitoring, J. Clin. Oncology 11: 1189-
1197.
Wedderburn, R.M.W. (1974). Quasilikelihood functions, generalized linear
models and the Gauss-Newton method, Biometrika 61: 439-447.
Wei, L.J. and Stram, D.O. (1988). Analying repeated measurements with
possibly missing observations by modelling marginal distributions,
Statist. Med. 7:139-148.
White, H. (1980). A heteroskedasticity consistent covariance matrix estima-
tor and a direct test for heteroskedasticity, Econometrica 48: 817-850.
White, H. (1982). Maximum likelihood estimation of misspecified models,
Econometrica 50: 1-25.
Whitehead, J . (1986). Sample sizes for phase I1 and phase I11 clinical trials,
in integrated approach, Statist. Med. 5: 459-464.
Whitehead, J. (1992). The Design and Analysis of Sequential Clinical
Trials, 2nd ed., Ellis Harwood, New York.
Whitehead, J . (1993). The case for frequentism in clinical trials, Statist.
Med. 12: 1405-1415.
Whitehead, J. and Brunier, H. (1995). Bayesian decision procedures for
dose determining experiments, Statist. Med. 14: 885-893.
Wild, C.J. and Yee, T.W. (1996). Additive extensions to generalized esti-
mating equations methods, J. Roy. Statist. SOC.B58: 711-725.
Williams, D.A. (1987). Generalized linear model diagnostics using the
deviance and single case deletions, Appl. Statist. 36: 181-191.
Wittes, J . and Wallenstein, S. (1987). The power of the Mantel-Haenszel
test, J. A m . Statist. Assoc. 82: 1104-1109.
Wingen, A.M., FabianBach, C., Schaefer, F. and Mehls, 0. (1997). Ran-
domized multicentre study of low-protein diet on the progression of
chronic renal failure in children, Lancet 349: 1117-1123.
References 331

Yates, F. (1982). Regression models for repeated measurements, Biometrics


38: 850-853.
Yusuf, S., Willes, J., Probstfield, J. and Tyroler, H.A. (1991). Analysis and
interpretation of treatment effects in subgroups of patients in random-
ized cIinica1 trials, J. Am. Med. Assoc. 266: 93-98.
Zeger, S.L. and Ka.tz, J. (1998). Graphical presentat,ion of longitudinal
data. In Encyclopedia of Biostatistics (eds. P. Armitage and T.
Colton), Wiley, Chichester.
Zeger, S.L., Liang, K.-Y. and Albert, P.S. (1988). Models for longitudi-
nal data: a generalized estimating equation approach, Biornetrics 44:
1049-1060.
Zeger, S.L and Liang, K-Y. (1986). Longitudinal dat.a analysis for discrete
and continuous outcomes, Biometrics 42: 121-130.
Zeger, S.L. and Qaqish, B. (1992). Multivariate regression analyses for
categorical data (with discussion), J. Roy. Statist. SOC.€354: 3-40.
Zeger, S.L., Diggle, P.J. and Huang, W. (1998). Generalized linear mod-
els for longitudinal data. In Encyclopedia of Biostatisstics (eds. P.
Armitage and T. Cotton), Wiley, Chichester.
Zeien, M. (1983). Guidelines for publishing papers on cancer clinical trials:
responsibilities of editors and authors, J. Clin. Oncol. 1: 164-169.
Zhao, L.P. and Prentice, R.L. (1990). Correlated binary regression using a
quadratic exponential model, Biometrika 77:642-648.
Index
a quality of life assessment, 2 Bonferroni procedure, 94
absolute risk reduction bootstrap, 181
(ARR), 252 bootstrap estimation, 102
accelerated failure time bootstrap resampling, 100
models, 221 boxplot, 71
acceleration factor, 222 BUGS, 309
adaptive randomisation burn-in, 260
procedure, 26
adherers-only method, 45
capsule count, 44
age specific failure rate, 219
censored observations, 208
alpha spending procedure, 58
change, 130
analysis of covariance, 24
clinical trial, 1
analysis-as-randomised, 45
ANC0L7A, 130 clinical trials, 16
aut,oregressive, 186 placebo-controlled. 6
aut,oregressive correlation cluster randomisation, 35
matrix: 184 cluster sampling, 247
auxiliary variables, 249 competing risks. 248
AZT as a therapy for AIDS: 13 compliance, 41
compound symmetry, 153
Baseline measurements, 129 confidence interval, 292
bath tub shape, 220 CONSORT. 284
Bayes factors, 263 continuation-ratio model, 89
Bayes’ theorem, 66, 256 control treatment, 1
Bayesian statistics, 255 Cook-Heisberg test, 190
bias-corrected percentile, 105 correlated response variables, 178
biased estimators of, 151 cost-benefit analysis, 101
binary variables, 108 covariate-dependent dropout, 52
binomial data, 141 Cox model, 226
blinding, 13 Cox-Snell residual, 235
blocked randomisation, 22 cross-over design, 6
BMDPSV, 176, 307 cumulative hazard function, 219

333
334 Index

decision theory, 275 Gaussian quadrature, 196


deletion diagnostics, 242 generalised estimating equations
delta-beta (Ap), 85 (GEE), 183
deviance residual, 235 generalised linear model, 77
dose-escalation experiments, 7 GENSTAT, 307
double-blind, 14 GLMM, 307
dropout, 45 GLMs, 78
ignorable, 147 gold-standard, 14
Missing at randum (MAR), Goodness-of-fit tests, 240
52 group sequential, 55
Missing completely at group sequential procedures, 65
randum (MCAR), 52
Non-ignorable (informative) , hazard function, 209, 219
52 Hessian matrix, 78
heteroscedastic consistent, 81
EaSt, 306 histogram, 71
East manual, 59 historical controls , 2
economic evaluations, 97 hotelling’s T 2 ,95
effect size, 292
efficient score residuals, 236 independence working model, 180
EGRET, 307 individual and collective
electronic monitoring, 44 ethics, 13
EM algorithm, 174 influence, 86, 241
ethics of clinical informative priors, 268
instantaneous failure rate, 219
trials, 9
integrated hazard function, 219
exchangeable, 186
exchangeable correlation intensity function, 219
matrix, 184 intention-to-treat, 45
exponential distribution, 2 10 interaction, 6
interim analysis, 24
factorial design, 6 inverse-Gaussian model. 89
familywise error (FWE), 94
James Lind, 2
finite mixture model, 174
Food and Drug Administration Kaplan-Meier, 213
(FDA), 40 Karnofsky Performance
frailty models, 250 Index, 223
full probability model, 257
funnel plot, 284 last observation carried forward
(LOCF), 125
gain function, 275 link function, 79
Index 335

log-rank, 215 non-stationarity, 146


logistic regression, 77 nonlinear models, 128
longitudinal data, 108 nQuery Advisor Version 2.0, 305
nuisance parameters, 144
Mahalanobis distances, 159 number needed to treat, 253
MANOVA, 168 number of deaths avoided, 253
Mantel-Haenszel test, 215
marginal models, 179 odds ratio, 73
marginal opportunity cost, 98 OSWALD, 169
marginal, random effects, 179 outcome measure, 2
Markov Chain Monte overdispersion, 82
Carlo (MCMC), 259
Martingale residual, 235 parallel groups design, 5
Matching, 246 partial score residuals, 239
maximum likelihood patient compliance, 38
estimates, 146 Pearson residuals, 82
measuring costs, 97 percentile method, 105
median survival time, 215 permuted block
meta-analysis, 31, 280 randomisation, 22
minimisation, 26 PEST3, 306
Missing at random (MAR), Phase I trials, 7
see dropout Phase I1 trials, 7
Missing completely random Phase I11 trials, 8
(MCAR), see dropout Phase IV trials, 8
missing values, 51 pill, 44
mixed effects model, 153 play the winner rules, 35
MLWiN, 173 Poisson regression, 78
monotone data pattern, 51 populat ion-averaged
multicentre trials, 8 parameter, 180
multiple endpoints, 36 POST, 130
multiple regression, 158 posterior distribution, 66, 256
multiple testing, 55 power, 26
multivariate normal distribution, power curves, 132
166 prior distribution, 66, 256
Mx, 173 product limit estimator, 213
prognostic factors, 24
non-complier, 41 proportional hazards regression
non-normal outcome model, 221
variables, 108 proportional-odds model, 89
non-parallel models, 90 protocol, 16
336 Index

quasi-score functions, 82 structural equation


quasilikelihood, 82 modelling, 173
subgroup analysis, 34
random allocation, 5 subject-specific intercepts, 194
random coefficients model, 161 SUDAAN, 308
random effects structure, 166 summary measures, 120, 121
randomisation, 2, 4 surrogate variable, 40
range of equivalence, 269 survival analysis, 81
regression to the mean, 130 survival data, 208
reliability, 40 survivor function, 209
residuals, 158
response feature, 120 taxonomy of clinical trials, 7
restricted maximum taxonomy of dropouts, 50
likelihood (REML), 151 time-by-time
right-censoring, 209 analysis, 118
robust, 81 time-varying covariates, 242
tracking, 112
sandwich estimator, 81 transition, 179
SAS, 176, 306 transition models, 197
SAS PROC MIXED, 307 treatment assignment , 4
scatterplot matrix, 114 treatment-received method, 45
Schoenfeld residuals, 236 two-sample t-test, 72
score equations, 78 type I error, 29
shrinkage, 195 type I1 error rate, 31
side-by-side box plots, 112
size of a clinical trial, 29 unbalanced design, 28
simple randomisation, 21 unequal group sizes, 22
single case deletion diagnostic, 86 unequal randomisation, 27
single-blind , 14 unstructured, 154, 186
spikeplot, 83
SPSS, 307 variance function, 80
STATA, 307
stratification, 245 Wald’s test, 167
stratified analysis, 26 Weibull distribution, 210
stratified randomisation, 24 Wilcoxon rank-sum test, 72

You might also like